16 research outputs found

    Spin Selectivity in Chiral Linked Systems

    Full text link
    "This is the peer reviewed version of the following article: Ageeva, Aleksandra A., Ekaterina A. Khramtsova, Ilya M. Magin, Denis A. Rychkov, Peter A. Purtov, Miguel A. Miranda, and Tatyana V. Leshina. 2018. "Spin Selectivity in Chiral Linked Systems." Chemistry - A European Journal 24 (15). Wiley: 3882-92. doi:10.1002/chem.201705863, which has been published in final form at https://doi.org/10.1002/chem.201705863. This article may be used for non-commercial purposes in accordance with Wiley Terms and Conditions for Self-Archiving."[EN] This work has shown spin selectivity in electron transfer (ET) of diastereomers of (R,S)-naproxen-(S)-N-methylpyrrolidine and (R,S)-naproxen-(S)-tryptophan dyads. Photoinduced ET in these dyads is interesting because of the still unexplained phenomenon of stereoselectivity in the drug activity of enantiomers. The chemically induced dynamic nuclear polarization (CIDNP) enhancement coefficients of (R,S)-diastereomers are double those of the (S,S)-analogue. These facts are also interesting because spin effects are among the most sensitive, even to small changes in spin and molecular dynamics of paramagnetic particles. Therefore, CIDNP reflects the difference in magnetoresonance parameters (hyperfine interaction constants (HFIs), g-factor difference) and lifetimes of the paramagnetic forms of (R,S)- and (S,S)-diastereomers. The difference in HFI values for diastereomers has been confirmed by a comparison of CIDNP experimental enhancement coefficients with those calculated. Additionally, the dependence of the CIDNP enhancement coefficients on diastereomer concentration has been observed for the naproxen-N-methylpyrrolidine dyad. This has been explained by the participation of ET in homo-(R,S-R,S or S,S-S,S) and hetero-(R,S-S,S) dimers of dyads. In this case, the effectivity of ET, and consequently, CIDNP, is supposed to be different for (R,S)- and (S,S)-homodimers, heterodimers, and monomers. The possibility of dyad dimer formation has been demonstrated by using high-resolution X-ray and NMR spectroscopy techniques.The work was supported by the Russian Foundation for Fundamental Research (14-03-00192).Ageeva, A.; Khramtsova, E.; Magin, I.; Richkov, D.; Purtov, P.; Miranda Alonso, MÁ.; Leshina, T. (2018). Spin Selectivity in Chiral Linked Systems. Chemistry - A European Journal. 24(15):3882-3892. https://doi.org/10.1002/chem.201705863S388238922415Lin, G.-Q., Zhang, J.-G., & Cheng, J.-F. (2011). Overview of Chirality and Chiral Drugs. Chiral Drugs, 3-28. doi:10.1002/9781118075647.ch1Lin, G.-Q., You, Q.-D., & Cheng, J.-F. (Eds.). (2011). Chiral Drugs. doi:10.1002/9781118075647Krasulova, K., Siller, M., Holas, O., Dvorak, Z., & Anzenbacher, P. (2015). Enantiospecific effects of chiral drugs on cytochrome P450 inhibitionin vitro. Xenobiotica, 46(4), 315-324. doi:10.3109/00498254.2015.1076086Shen, Q., Wang, L., Zhou, H., Jiang, H., Yu, L., & Zeng, S. (2013). Stereoselective binding of chiral drugs to plasma proteins. Acta Pharmacologica Sinica, 34(8), 998-1006. doi:10.1038/aps.2013.78Duggan, K. C., Hermanson, D. J., Musee, J., Prusakiewicz, J. J., Scheib, J. L., Carter, B. D., 
 Marnett, L. J. (2011). (R)-Profens are substrate-selective inhibitors of endocannabinoid oxygenation by COX-2. Nature Chemical Biology, 7(11), 803-809. doi:10.1038/nchembio.663JimĂ©nez, M. C., Pischel, U., & Miranda, M. A. (2007). Photoinduced processes in naproxen-based chiral dyads. Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 8(3), 128-142. doi:10.1016/j.jphotochemrev.2007.10.001Khramtsova, E. A., Sosnovsky, D. V., Ageeva, A. A., Nuin, E., Marin, M. L., Purtov, P. A., 
 Leshina, T. V. (2016). Impact of chirality on the photoinduced charge transfer in linked systems containing naproxen enantiomers. Physical Chemistry Chemical Physics, 18(18), 12733-12741. doi:10.1039/c5cp07305gKhramtsova, E. A., Ageeva, A. A., Stepanov, A. A., Plyusnin, V. F., & Leshina, T. V. (2017). Photoinduced Electron Transfer in Dyads with (R)-/(S)-Naproxen and (S)-Tryptophan. Zeitschrift fĂŒr Physikalische Chemie, 231(3). doi:10.1515/zpch-2016-0842Abad, S., Pischel, U., & Miranda, M. A. (2005). Intramolecular electron transfer in diastereomeric naphthalene–amine dyads: a fluorescence and laser flash photolysis study. Photochem. Photobiol. Sci., 4(1), 69-74. doi:10.1039/b409729gLevkin, P. A., Kokorin, A. I., Schurig, V., & Kostyanovsky, R. G. (2006). Solid-state ESR differentiation between racemate versus enantiomer. Chirality, 18(4), 232-238. doi:10.1002/chir.20242Khlestkin, V. K., Glasachev, Y. I., Kokorin, A. I., & Kostyanovsky, R. G. (2004). ESR study of stereochemistry in chiral nitroxide radical crystals. Mendeleev Communications, 14(6), 318-320. doi:10.1070/mc2004v014n06abeh002055MĂ€urer, M., & Stegmann, H. B. (1990). Chiral recognition of diastereomeric esters and acetals by EPR and NMR investigations. Chemische Berichte, 123(8), 1679-1685. doi:10.1002/cber.19901230817Kreilick, R. W., Becher, J., & Ullman, E. F. (1969). Stable free radicals. V. Electron spin resonance studies of nitronylnitroxide radicals with asymmetric centers. Journal of the American Chemical Society, 91(18), 5121-5124. doi:10.1021/ja01046a032Schuler, P., Schaber, F.-M., Stegmann, H. B., & Janzen, E. (1999). Recognition of chirality in nitroxides using EPR and ENDOR spectroscopy. Magnetic Resonance in Chemistry, 37(11), 805-813. doi:10.1002/(sici)1097-458x(199911)37:113.0.co;2-kDoktorov, A. B., Mikhailov, S. A., & Purtov, P. A. (1992). Theory of geminate recombination of radical pairs with instantaneously changing spin-Hamiltonian. I. General theory and kinematic approximation. Chemical Physics, 160(2), 223-237. doi:10.1016/0301-0104(92)80124-eMagin, I. M., Purtov, P. A., Kruppa, A. I., & Leshina, T. V. (2005). Peculiarities of Magnetic and Spin Effects in a Biradical/Stable Radical Complex (Three-Spin System). Theory and Comparison with Experiment. The Journal of Physical Chemistry A, 109(33), 7396-7401. doi:10.1021/jp051115yYin, P., Zhang, Z.-M., Lv, H., Li, T., Haso, F., Hu, L., 
 Liu, T. (2015). Chiral recognition and selection during the self-assembly process of protein-mimic macroanions. Nature Communications, 6(1). doi:10.1038/ncomms7475Soai, K., Kawasaki, T., & Matsumoto, A. (2014). The Origins of Homochirality Examined by Using Asymmetric Autocatalysis. The Chemical Record, 14(1), 70-83. doi:10.1002/tcr.201300028Frank, F. C. (1953). On spontaneous asymmetric synthesis. Biochimica et Biophysica Acta, 11, 459-463. doi:10.1016/0006-3002(53)90082-1Hegstrom, R. A., & Kondepudi, D. K. (1996). Influence of static magnetic fields on chirally autocatalytic radical-pair reactions. Chemical Physics Letters, 253(3-4), 322-326. doi:10.1016/0009-2614(96)00248-5Ishida, Y., & Aida, T. (2002). Homochiral Supramolecular Polymerization of an «S»-Shaped Chiral Monomer:  Translation of Optical Purity into Molecular Weight Distribution. Journal of the American Chemical Society, 124(47), 14017-14019. doi:10.1021/ja028403hSato, K., Itoh, Y., & Aida, T. (2014). Homochiral supramolecular polymerization of bowl-shaped chiral macrocycles in solution. Chem. Sci., 5(1), 136-140. doi:10.1039/c3sc52449cDubinets, N. O., Safonov, A. A., & Bagaturyants, A. A. (2016). Structures and Binding Energies of the Naphthalene Dimer in Its Ground and Excited States. The Journal of Physical Chemistry A, 120(17), 2779-2782. doi:10.1021/acs.jpca.6b03761‘Excimer’ fluorescence VII. Spectral studies of naphthalene and its derivatives. (1965). Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 284(1399), 551-565. doi:10.1098/rspa.1965.0080Kerr, H. E., Softley, L. K., Suresh, K., Hodgkinson, P., & Evans, I. R. (2017). Structure and physicochemical characterization of a naproxen–picolinamide cocrystal. Acta Crystallographica Section C Structural Chemistry, 73(3), 168-175. doi:10.1107/s2053229616011980Saprygina, N. N., Morozova, O. B., Gritsan, N. P., Fedorova, O. S., & Yurkovskaya, A. V. (2011). 1H CIDNP study of the kinetics and mechanism of the reversible photoinduced oxidation of tryptophyl-tryptophan dipeptide in aqueous solutions. Russian Chemical Bulletin, 60(12), 2579-2587. doi:10.1007/s11172-011-0396-0Schwarz, W., Dangel, K. M., Bargon, J., & Jones, G. (1982). CIDNP studies of photoinitiated electron-transfer reactions. Sensitized isomerization of an electron-acceptor norbornadiene. Journal of the American Chemical Society, 104(21), 5686-5689. doi:10.1021/ja00385a022Gilbert, B. C., Larkin, J. P., & Norman, R. O. C. (1972). Electron spin resonance studies. Part XXXIV. The use of the aci-anion from nitromethane as a spin trap for organic radicals in aqueous solution. Journal of the Chemical Society, Perkin Transactions 2, (9), 1272. doi:10.1039/p29720001272MĂ€urer, M., Stegmann, H. B., Hiller, W., & MĂŒller, B. (1992). Stereoelectronic and Steric Effects in the Synthesis and Recognition of Diastereomeric Ethers by NMR and EPR Spectroscopy. Chemische Berichte, 125(4), 857-865. doi:10.1002/cber.19921250417Pejov, L. (2001). A gradient-corrected density functional study of indole self-association through N–Hâ‹ŻÏ€ hydrogen bonding. Chemical Physics Letters, 339(3-4), 269-278. doi:10.1016/s0009-2614(01)00341-4Muñoz, M. ., Ferrero, R., Carmona, C., & BalĂłn, M. (2004). Hydrogen bonding interactions between indole and benzenoid-π-bases. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 60(1-2), 193-200. doi:10.1016/s1386-1425(03)00206-3Hatton, J. V., & Richards, R. E. (1962). Solvent effects in N.M.R. spectra of amide solutions. Molecular Physics, 5(2), 139-152. doi:10.1080/00268976200100141VayĂĄ, I., Andreu, I., JimĂ©nez, M. C., & Miranda, M. A. (2014). Photooxygenation mechanisms in naproxen–amino acid linked systems. Photochem. Photobiol. Sci., 13(2), 224-230. doi:10.1039/c3pp50252jGavezzotti, A. (1994). Are Crystal Structures Predictable? Accounts of Chemical Research, 27(10), 309-314. doi:10.1021/ar00046a004Gavezzotti, A., & Filippini, G. (1994). Geometry of the Intermolecular X-H.cntdot..cntdot..cntdot.Y (X, Y = N, O) Hydrogen Bond and the Calibration of Empirical Hydrogen-Bond Potentials. The Journal of Physical Chemistry, 98(18), 4831-4837. doi:10.1021/j100069a010Closs, G. L., & Miller, R. J. (1979). Laser flash photolysis with NMR detection. Microsecond time-resolved CIDNP: separation of geminate and random-phase processes. Journal of the American Chemical Society, 101(6), 1639-1641. doi:10.1021/ja00500a068Goez, M. (1992). Pseudo steady-state photo-CIDNP measurements. Chemical Physics Letters, 188(5-6), 451-456. doi:10.1016/0009-2614(92)80847-5Rychkov, D. A., Arkhipov, S. G., & Boldyreva, E. V. (2014). Simple and efficient modifications of well known techniques for reliable growth of high-quality crystals of small bioorganic molecules. Journal of Applied Crystallography, 47(4), 1435-1442. doi:10.1107/s160057671401127

    Effect of Amino Group Charge on the Photooxidation Kinetics of Aromatic Amino Acids

    No full text
    The kinetics of the photooxidation of aromatic amino acids histidine (His), tyrosine (Tyr), and tryptophan (Trp) by 3,3â€Č,4,4â€Č-benzophenonetetracarboxylic acid (TCBP) has been investigated in aqueous solutions using time-resolved laser flash photolysis and time-resolved chemically induced dynamic nuclear polarization. The pH dependence of quenching rate constants is measured within a large pH range. The chemical reactivities of free His, Trp, and Tyr and of their acetylated derivatives, <i>N</i>-AcHis, <i>N</i>-AcTyr, and <i>N</i>-AcTrp, toward TCBP triplets are compared to reveal the influence of amino group charge on the oxidation of aromatic amino acids. The bimolecular rate constants of quenching reactions between the triplet-excited TCBP in the fully deprotonated state and tryptophan, histidine, and tyrosine with a positively charged amino group are <i>k</i><sub>q</sub> = 2.2 × 10<sup>9</sup> M<sup>–1</sup> s<sup>–1</sup> (4.9 < pH < 9.4), <i>k</i><sub>q</sub> = 1.6 × 10<sup>9</sup> M<sup>–1</sup> s<sup>–1</sup> (6.0 < pH < 9.2), and <i>k</i><sub>q</sub> = 1.5 × 10<sup>9</sup> M<sup>–1</sup> s<sup>–1</sup> (4.9 < pH < 9.0), respectively. Tryptophan, histidine, and tyrosine with a neutral amino group quench the TCBP triplets with the corresponding rate constants <i>k</i><sub>q</sub> = 8.0 × 10<sup>8</sup> M<sup>–1</sup> s<sup>–1</sup> (pH > 9.4), <i>k</i><sub>q</sub> = 3.0 × 10<sup>8</sup> M<sup>–1</sup> s<sup>–1</sup> (pH > 9.2), and <i>k</i><sub>q</sub> = (4.0–10.0) × 10<sup>8</sup> M<sup>–1</sup> s<sup>–1</sup> (9.0 < pH < 10.1) that are close to those for the N-acetylated derivatives. Thus, it has been established that the presence of charged amino group changes oxidation rates by a significant factor; i.e., His with a positively charged amino group quenches the TCBP triplets 5 times more effectively than <i>N</i>-AcHis and His with a neutral amino group. The efficiency of quenching reaction between the TCBP triplets and Tyr and Trp with a positively charged amino group is about 3 times as high as that of both Tyr and Trp with a neutral amino group, <i>N</i>-AcTyr and <i>N</i>-AcTrp
    corecore