Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-29T21:43:41.446Z Has data issue: false hasContentIssue false

Origin of the homochirality of biomolecules

Published online by Cambridge University Press:  17 March 2009

L. Keszthelyi
Affiliation:
Institute of Biophysics, Biological Research Centre, Szeged, Hungary, H-6701.

Extract

Molecules built up from a given set of atoms may differ in their three-dimensional structure. They may have one or more asymmetric centres that serve as reference points for the steric distribution of the atoms. Carbon atoms, common to all biomolecules, are often such centres. For example, the Cα atom between the carboxyl and amino groups in amino acids is an asymmetric centre: looking ON ward (i.e. from the carbOxyl to the amiNo group, with the Cα oriented so that it is above the carboxyl and amino groups) the radical characterizing the amino acid may be to the right (D-molecules) or to the left (L-molecules). Nineteen of the 20 amino acids occurring in proteins have such a structure (the exception is glycine, where the radical is a hydrogen atom). These pairs of molecules cannot be brought into coincidence with their own mirror image, as is the situation with our hands. The phenomenon has therefore been named handedness, or chirality, from the Greek word cheir, meaning hand. The two forms of the chiral molecules are called enantiomers or antipodes. They differ in rotating the plane of the polarized light either to the right or to the left. The sense of rotation depends on the wavelength of the measuring light, but at a given wavelength it is always opposite for a pair of enantiomers. Chirality may also occur when achiral molecules form chiral substances during crystallization (for example, quartz forms D-quartz or Lquartz). A detailed theoretical treatment of molecular chirality is given by Barron (1991).

Type
Research Article
Copyright
Copyright © Cambridge University Press 1995

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Agarwal, V. K., Schutte, W., Greenberg, J. M., Ferris, J. P., Briggs, R., Connor, S., de Bult, Van, C. P. E. M., & Baas, F. (1985). Photochemical reactions in interstellar grains photolysis of CO, NH3, and H2O. Origins of Life 16, 2140.Google Scholar
Ageno, M. (1972). On the molecular asymmetry of living organisms. J. Theor. Biol. 37, 187192.Google Scholar
Angel, G. P. R., Illing, R. & Martin, P. G. (1972). Circular polarization of the twilight. Nature 238, 389390.Google Scholar
Avetisov, V. A. & Goldanskii, V. I. (1991). Homochirality and stereospecific activity: evolutionary aspects. BioSystems 25, 141149.CrossRefGoogle ScholarPubMed
Avetisov, V. A., Goldanskii, V. I. & Kuz'min, V. V. (1991). Handedness, origin of life and evolution. Physics Today, July, 3341.CrossRefGoogle ScholarPubMed
Avetisov, V. A., Kuz'min, V. V. & Anikin, S. A. (1987). Sensitivity of chemical chiral systems to weak asymmetric factors. Chem. Phys. 112, 179187.CrossRefGoogle Scholar
Bada, J. L., Cronin, J. R., Ho, M. S., Kvenvolden, K. A., Lawless, J. G., Miller, S. L., Oro, J. & Steinberg, S. (1983). On the reported optical activity of amino acids in the Murchison meteorite. Nature, 301, 494496. See also the reply of Engel & Nagy, Nature, 301, 496–497.Google Scholar
Bada, J. L. & Miller, S. L. (1987). Racemization and the origin of optically active organic compounds in living organisms. BioSystems 20, 2126.CrossRefGoogle ScholarPubMed
Barron, L. D. (1986). True and false chirality and parity violation. Chem. Phys. Lett. 123, 423427.CrossRefGoogle Scholar
Barron, L. D. (1991). Fundamental symmetry aspects of molecular chirality. In New Developments in Molecular Chirality (ed. Mezey, P. G.), pp. 155. The Netherlands: Kluwer Academic Publishers.Google Scholar
Barron, L. D. (1993). True and false chirality. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 267283. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Barron, L. D. (1994). Can a magnetic field induce absolute asymmetric synthesis? Science 266, 14911492.CrossRefGoogle ScholarPubMed
Beerlage, M. J. M., Farago, P. S. & der, VanWiel, M. J. (1981). A search for spin effects in low-energy electron scattering from optically active camphor. J. Phys. B. At. Mol. Phys. 14, 32463253.CrossRefGoogle Scholar
Bingham, D. K. & Evans, M. E. (1975). Precambrian geomagnetic field reversal. Nature 253, 332333.CrossRefGoogle Scholar
Bonner, W. A. (1991). The origin and amplification of biomolecular chirality. Origins of Life and Evolution of the Biosphere 21, 59111.CrossRefGoogle ScholarPubMed
Bonner, W. A. & Rubenstein, E. (1987). Supernovae, neutron stars and biomolecular chirality. BioSystems 20, 99111.Google Scholar
Brownlee, D. E. & Kissel, J. (1991). The composition of dust particles in the environment of Comet Halley. In Comet Halley, Vol. 2 (ed. Mason, J.), pp. 8998. New York, Toronto, Sidney, Tokyo: Ellis Horwood.Google Scholar
Campbell, D. M. & Farago, P. S. (1985). Spin-dependent electron scattering from optically active molecules. Nature 318, 5253.Google Scholar
Campbell, D. M. & Farago, P. S. (1987). Electron optic dichroism in camphor. J. Phys. B. At. Mol. Phys. 20, 51335143.CrossRefGoogle Scholar
Chela-Flores, J. (1991). Comments on a novel approach to the role of chirality in the origin of life. Chirality 3, 389392.Google Scholar
Chela-Flores, J. (1994). The origin of chirality in protein amino acids. Chirality 6, 165168.CrossRefGoogle Scholar
Chyba, C. F., Thomas, P. J., Brookshaw, L. & Sagan, C. (1990). Cometary delivery of Organic molecules to the early Earth. Science 249, 366373.Google Scholar
Czégé, J. & Fajszi, Cs. (1977). Remark on models for amplification of asymmetry. Origins of Life 8, 271275.Google Scholar
Diener, T. O. (1993). Small pathogenic RNAs of plants: living fossils of the RNA world? In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 6984. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Doty, P. & Lundberg, R. D. (1956). Polypeptides. X. Configurational and stereochemical effects in the amine initiated polymerization of N-carbon-anhydrides. J.Am. Chem. Soc. 78, 48104812.CrossRefGoogle Scholar
Dougherty, R. C. (1980). Chemical geometrodynamics: gravitational fields can influence the course of prochiral chemical reactions. J. Am. Chem. Soc. 102, 380381.CrossRefGoogle Scholar
Edwards, D., Cooper, K. & Dougherty, R. C. (1980). Asymmetric synthesis in a confined vortex: gravitational fields can cause asymmetry, J. Am. Chem. Soc. 102, 381382.CrossRefGoogle Scholar
Eigen, M. & Schuster, P. (1979). The Hypercycle. Berlin, Heidelberg, New York: Springer.CrossRefGoogle Scholar
Engel, M. H., Macro, S. A. & Silfer, J. A. (1990). Carbon isotope composition of individual amino acids in the Murchison meteorite. Nature 348, 4749.CrossRefGoogle ScholarPubMed
Engel, M. H. & Nagy, B. (1982). Distribution and enantiomeric composition of amino acids in the Murchison meteorite. Nature 296, 837840.CrossRefGoogle Scholar
Epstein, S., Krishnamurty, R. V., Cronin, J. R., Pizzarello, S. & Yuen, G. U. (1987). Unusual stable isotope ratios in amino acid and carboxylic acid extracts from the Murchison meteorite. Nature 326, 477479.CrossRefGoogle ScholarPubMed
Fajszi, Cs. & Czégé, J. (1977). β-decay and the origin of optical purity. Origins of Life 8, 277281.CrossRefGoogle ScholarPubMed
Feynman, R. P., Leighton, R. B. & Sands, M. (1964 a). Feynman's Lectures on Physics, Vol. I, pp. 52. 1–12. Reading, Massachusetts, Palo Alto, London: Addison-Wesley Publishing Company, Inc.Google Scholar
Feynman, R. P., Leighton, R. B. & Sands, M. (1964 b). Feynman's Lectures on Physics, Vol. II, pp. 61. 9–11. Reading, Massachusetts, Palo Alto, London: Addison-Wesley Publishing Company Inc.Google Scholar
Figureau, A., Duval, E. & Boukenter, A. (1993). Search for phase transitions changing molecular chirality. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 157164. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Garay, A. S. (1993). Theoretical and experimental studies of the possibility of chirality dependent time direction in molecules. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 165179. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Garay, A. S. & Ahlgren-Beckendorf, J. A. (1990). Differential interaction of chiral β-particles with enantiomers. Nature 346, 451453.Google Scholar
Garay, A. S., Ahlgren-Beckendorf, J. A. & Gazdy, B. (1994). Molecular handedness and chiral strength determined by matter-wave circular dichroism. Phys. Rev. A 48, 30083011.CrossRefGoogle Scholar
Garay, A. S., Keszthelyi, L., Demeter, I. & Hrasko, P. (1974). Origin of asymmetry in biomolecules. Nature 250, 332333.CrossRefGoogle ScholarPubMed
Gay, T. J., Trenthem, K. W. & Johnston, M. L. (1995). Lecture at the Peter Farago Symposium, 31 March-1 Apr. Edinburgh, to be published in the Proceedings. New York: Plenum Press.Google Scholar
Gerike, P. (1975). A new absolute asymmetric synthesis. Naturwiss. 62, 3839.Google Scholar
Gidley, D. W., Rich, A., Van House, J. & Zitzewitz, P. W. (1985). β decay and the origin of biological chirality: experimental results. Nature 297, 639643.CrossRefGoogle Scholar
Gilat, G. (1985). Chiral interactions in biomolecules. Chem. Phys. Lett. 121, 913.CrossRefGoogle Scholar
Gilat, G. (1993). Chiral interaction and biomolecular evolution. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 285293. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Gould, S. J. (1994). The evolution of life on the Earth. Sci. Am. 271, (10) 6269.Google Scholar
Greenberg, J. M. (1984). The structure and evolution of interstellar grains. Sci. Am. 250, (6) 124135.CrossRefGoogle Scholar
Hegstrom, R. A. (1982). β decay and the origins of biological chirality: theoretical results. Nature 297, 643647.CrossRefGoogle ScholarPubMed
Hegstrom, R. A., Rich, A. & Van House, J. (1985). New estimates of asymmetric decomposition of racemic mixtures by natural β-radiation sources. Nature 313, 391392.CrossRefGoogle ScholarPubMed
Hochstim, A. R. (1975). Non-linear mathematical models for the origin of asymmetry in biological molecules. Origins of Life 6, 317366.CrossRefGoogle ScholarPubMed
Jelley, J. F. (1958). Cerenkov Radiation and its Application. London, New York: Pergamon.Google Scholar
Joyce, G. F. (1989). RNA evolution and the origin of life. Nature 338, 217224.Google Scholar
Joyce, G. F., Visser, G. M., van Boeckel, C. A. A., van Boom, J. H.Orgel, L. E. & van Westrenen, J. (1984). Chiral selection in poly(C)-directed synthesis of oligo(G). Nature 310, 602604.CrossRefGoogle ScholarPubMed
Kessler, J.Lecture at the Peter Farago Symposium, 31 March-1 Apr. Edinburgh, to be published in the Proceedings. New York: Plenum Press.Google Scholar
Keszthelyi, L. (1976). Chemical evolution: effect of high energy radiation. Origins of Life 7, 349354.CrossRefGoogle ScholarPubMed
Keszthelyi, L. (1977 a). Contribution of parity-violating effects to intramolecular interactions. Phys. Lett. 64A, 287288.CrossRefGoogle Scholar
Keszthelyi, L. (1977 b). Origin of the asymmetry of biomolecules and weak interaction. Origins of Life 8, 299340.CrossRefGoogle ScholarPubMed
Keszthelyi, L. (1978). Origin of the asymmetry of biomolecules and the weak interaction: a critical review. In Origin of Life, Proceedings of the Second ISSOL Meeting (Ed. Noda, H.), pp. 327332. Japan: Center for Academic Publications.Google Scholar
Keszthelyi, L. (1981). Summary of round table discussion on status of research on ‘Generation and amplification of chirality in chemical systems’). Origins of Life 11, 191194.Google Scholar
Keszthelyi, L. (1984). Parity violation as a source of chirality in nature. Origins of Life 14, 378383.Google Scholar
Keszthelyi, L. (1995). Possibilities to measure the parity-violating energy difference. In Chemical Evolution: Structure and Model of the First Cell, (Ed. Ponnamperuma, C. and Chela-Flores, J.). The Netherlands, Kluwer Academic Publishers (In press).Google Scholar
Kondepudi, D. K., Bullock, R. L., Digits, J. A., Hall, J. K. & Miller, J. M. (1993). Kinetics of chiral symmetry breaking in crystallization. J. Am. Chem. Soc. 113, 1012110216.Google Scholar
Kondepudi, D. K., Kaufman, R. J. & Singh, N. (1990). Chiral symmetry breaking in sodium chlorate crystallization. Science 250, 975976.CrossRefGoogle Scholar
Kondepudi, D. K. & Nelson, G. W. (1983). Chiral symmetry breaking in nonequilibrium systems. Phys. Rev. Lett 50, 10231026.CrossRefGoogle Scholar
Kondepudi, D. K. & Nelson, G. W. (1985). Weak neutral currents and the origin of biomolecular chirality. Nature 314, 438441.CrossRefGoogle Scholar
Kovács, K. L. (1979). On the physical origin of biological handedness. Origins of Life 9, 219233.CrossRefGoogle ScholarPubMed
Kovács, K., Keszthelyi, L. & Goldanskii, V. J. (1981). Unconsidered sources of chirality in nature. Origins of Life 11, 93103.Google Scholar
Lee, T. D. & Yang, C. N. (1956). Question of parity conservation in weak interaction. Phys. Rev. 104, 254258.CrossRefGoogle Scholar
MacDermott, A. J. (1993). The weak force and the origin of life. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 85117. Hampton, Virginia: A. Deepak Publishing.Google Scholar
MacDermott, A. J. (1995 a). Search for extraterrestrial homochirality. In Chemical Evolution: Structure and Model of the First Cell (Ed. Ponnamperuma, C. and Chela-Flores, J.). The Netherlands: Kluwer Academic Publishers (In press).Google Scholar
MacDermott, A. J. (1995 b). Electroweak enantioselection and the origin of life. Origins of Life and Evolution of the Biosphere 25, 191199.CrossRefGoogle ScholarPubMed
MacDermott, A. J., Tranter, G. E. & Trainor, S. J. (1992). The search of large parity-violating energy differences finds fruit in thiosubstituted DNA analogues. Chem. Phys. Lett. 194, 152156.CrossRefGoogle Scholar
Marcus, J. N. & Olsen, M. A. (1991). Biological implications of organics in comets. In Comets in the Post-Halley Era (Ed. Newburn, R. L., Neugebauer, M. and Rahe, J.), Vol. I, pp. 439462. The Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Mason, S. F. (1983). Molecular handedness and the origins of chiral discrimination. Int. Rev. Phys. Chem. 3, 217241.Google Scholar
Mason, S. F. (1984). Origins of biomolecular handedness. Nature 311, 1923.Google Scholar
Mason, S. F. & Tranter, G. E. (1984). The parity-violating energy difference between enantiomeric molecules. Mol. Phys. 53, 10911111.CrossRefGoogle Scholar
McElhinny, M. W. (1971). Geomagnetic reversals during Phanerozoic. Science 172, 157159.CrossRefGoogle ScholarPubMed
Miller, S. L. (1953). A production of amino acids under possible primitive Earth conditions. Science 117, 528529.CrossRefGoogle ScholarPubMed
Morozov, L. (1979). Mirror symmetry breaking in biochemical evolution. Origins of Life 9, 187217.Google Scholar
Mörtberg, L. (1971). Nonbiotic origin of optical activity. Nature 232, 105107.CrossRefGoogle ScholarPubMed
Nagy, B. (1976). Organic chemistry of the young Earth. Naturwiss. 63, 499505.CrossRefGoogle Scholar
Nagy, B., Engel, M. H., Zumberge, J. E., Ogino, H. & Chang, S. Y. (1981). Amino acids and hydrocarbons ∼ 3,800-Myr old in the Isua Rocks, southwestern Greenland. Nature 289, 5356.CrossRefGoogle Scholar
Navarro-Gonzales, R., Khanna, R. K. & Ponnamperuma, C. (1993). Chirality and the origins of life. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 135155. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Norden, B. (1977). Was the photoresolution of amino acids the origin of optical activity in life? Nature 266, 567568.Google Scholar
Orgel, L. E. (1994). The origin of life on the Earth. Sci. Am. 271, (10) 5261.Google Scholar
Pigman, W. & Isbell, H. S. (1968). Mutarotation of sugars in solution. Part 1. Adv. Carbohydr. Chem. Biochem. 23, 1157.Google Scholar
Pincock, R. E., Perkins, R. R., Ma, A. S. & Wilson, K. R. (1971). Probability distribution of enantiomorphous forms in spontaneous generation of optically active substances. Science 174, 10181020.Google Scholar
Quack, M. (1986). On the measurement of the parity-violating energy difference between enantiomers. Chem. Phys. Lett. 132, 147153.CrossRefGoogle Scholar
Rein, D. (1993). Die wunderbare Händigkeit der Moleküle. Basel, Boston, Berlin: Birkhäuser.CrossRefGoogle Scholar
Rein, D. W., Hegstrom, R. A. & Sandars, P. G. H. (1979). Parity non-conserving energy difference between mirror image molecules. Phys. Lett. 71A, 499502.Google Scholar
Rutten, M. G. (1971). The Origin of Life by Natural Causes, pp. 206219. Amsterdam, London, New York: Elsevier.Google Scholar
Salam, A. (1991). The role of chirality in the origin of life. J. Mol. Evol. 33, 105113.Google Scholar
Salam, A. (1993). The origin of chirality, the role of phase transitions and their inductions in amino acids. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 101117. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Schopf, J. W. & Packer, B. M. (1992). Early archean (3·3-billion to 3·5-billion-yearold) microfossils from Warrawoona group, Australia. Science 237, 7073.Google Scholar
Schuster, P. (1993). Evolution in a RNA world. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 5168. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Schwartz, D. K. (1993). Instant patterns in thin films. Nature 362, 593594.Google Scholar
Shimoyama, A., Blair, N. & Ponnamperuma, C. (1978). In Origin of Life, Proceedings of the Second ISSOL Meeting (Ed. Noda, H.), pp. 9599. Japan: Center for Academic Publications.Google Scholar
Soós, J. & Bagyinka, Cs. (1976). On the differential β-decay in optical isomers. Radiobiochem. Radioanal. Lett. 27, 169175.Google Scholar
Stephen, T. M., Shi, X. & Burrow, P. D. (1988). Temporary negative-ion states of chiral molecules: camphor and 3-methylcyclopentanone. J. Phys. B. At. Mol. Phys. 21, L169L171.CrossRefGoogle Scholar
Tranter, G. E. (1985). Parity-violating energy differences of chiral minerals and the origin of biomolecular homochirality. Nature 318, 172173.Google Scholar
Ulbricht, T. L. V. (1962). The optical asymmetry of metabolites. In Comparative Biochemistry (Ed. Florkins, M. and Mason, H. S.), Vol. IV, pp. 125. New York: Academic Press.Google Scholar
Ulbricht, T. L. V. & Vester, F. (1962). Attempts to induce optical activity with polarized β-radiation. Tetrahedron 18, 629637.Google Scholar
Van House, J., Rich, A. & Zitzewitz, P. W. (1985). Limits of asymmetric orthopositronium formation in high Z optically active molecules. Origins of Life 16, 8190.Google Scholar
Vértes, A. & Nagy, D. L. (1990). Mössbauer Spectroscopy of Frozen Solutions. Budapest: Akadémiai Kiadó.Google Scholar
Wagniére, G. & Meier, A. (1983). Difference in the absorption coefficient of enantiomers for arbitrarily polarized light in a magnetic field: a possible source of chirality in molecular evolution. Experientia 39, 10901091.Google Scholar
Wald, G. (1957). The origin of optical activity. Ann. N. Y. Acad. Sci. 69, 352368.Google Scholar
Wang, W. Q., Wu, J. L. & Jang, J. (1993). Differences in radiolysis behavior of D, Lamino acid in primary stage and thermodynamic equilibrium state. In Chemical Evolution: Origin of Life (Ed. Ponnamperuma, C. and Chela-Flores, J.), pp. 249259. Hampton, Virginia: A. Deepak Publishing.Google Scholar
Wei-Min, L. (1982). Remarks on origin of biomolecular asymmetry. Origins of Life 12, 205209.CrossRefGoogle Scholar
Williams, K. M. & Smith, G. G. (1977). A critical evaluation of the application of amino acid racemization to geochronology and geothermometry. Origins of Life 8, 91144.CrossRefGoogle ScholarPubMed
Yamagata, Y. (1966). A hypothesis for the asymmetric appearance of biomolecules on Earth. J. Theor. Biol. 11, 495498.Google Scholar
Zadel, G., Eisenbraun, C., Wolff, G.-J. & Breitmeier, H. (1994). Enantiomeric selection in static magnetic field. Angew. Chem. Int. Ed. Engl. 33, 454460.Google Scholar