Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-28T21:40:37.591Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  17 December 2010

Pier Luigi Luisi
Affiliation:
ETH Zentrum, Switzerland
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
The Emergence of Life
From Chemical Origins to Synthetic Biology
, pp. 271 - 300
Publisher: Cambridge University Press
Print publication year: 2006

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

Abel, D. L. (2002). Is life reducible to complexity? In Fundamentals of Life, eds. Palyi, G., Zucchi, C., and Caglioti, L.. Elsevier, pp. 57–72.Google Scholar
Achilles, T. and Kiedrowski, G. (1993). A self-replicating system from three starting materialsAngew. Chem., 32, 1198–201.CrossRefGoogle Scholar
Alberts, B., Bray, D., Lewis, J., et al. (1989). Molecular Biology of the Cell, 2nd edn. Garland Publications Incorporate.Google Scholar
Alexander, S. (1920). Space, Time and Deity. McMillan.Google Scholar
Allison, A. C. and Gregoriadis, G. (1974). Liposomes as immunological adjuvants. Nature, 252, 252–8.CrossRefGoogle ScholarPubMed
Ambartsumian, T. G., Adamian, S. Y., Petrosia, L. S., and Simonian, A. L. (1992). Incorporation of water-soluble enzymes glucose-oxidase and urate oxidase into phosphatidylcholine liposomes. Biol. Membr., 5, 1878–87.Google Scholar
Anderson, G. and Luisi, P. L. (1979). Papain-induced oligomerization of alpha amino acid esters. Helv. Chim. Acta, 62, 488–94.CrossRefGoogle Scholar
Angelova, M. I. and Dimitrov, D. S. (1988). A mechanism of liposome electro-formation. Progr. Colloid Polymer. Sci., 76, 59–67.CrossRefGoogle Scholar
Annesini, M. C., Braguglia, C. M., Memoli, A., Palermiti, L. G., and Sario, Di S. (1997). Surfactant as modulating agent of enzyme-loaded liposome activity. Biotechnol. Bioeng., 55, 261–6.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Annesini, M. C., Giulio, Di A., Marzio, Di L., Finazzi-Agrò, A., and Mossa, G. (1992). J. Liposome Res., 2, 455–67.CrossRef
Annesini, M. C., Giorgio, Di L., Marzio, Di L., et al. (1993). J. Liposome Res., 3, 639–48.CrossRef
Annesini, M. C., Marzio, Di L., Finazzi-Agrò, A., Serafino, A. L., and Mossa, G. (1994). Interaction of cationic phospholipid-vesicles with carbonic anhydrase. Biochem. Mol. Biol. Int., 32, 87–94.Google ScholarPubMed
Apte, P. (2002). Vedantic view of Life. In Fundamentals of Life, eds. Palyi, G., Zucchi, C., and Caglioti, L.. Elsevier, pp. 497–502.Google Scholar
Arinin, E. I (2002). Essence of organic life in Russian orthodox and modern philosophical tradition: beyond functionalism and elementarism. In Fundamentals of Life, eds. Palyi, G., Zucchi, C., and Caglioti, L.. Elsevier, pp. 503–16.Google Scholar
Ashkenasy, G., Jagasia, R., Yadav, M., and Ghadiri, M. R. (2004). Design of a directed molecular network. Proc. Natl. Acad. Sci., 101, 10872–7.CrossRefGoogle ScholarPubMed
Atkins, P. and Paula, J. (2002). Physical Chemistry, 7th edn. Oxford University Press.Google Scholar
Atmanspacher, H. and Bishop, R. (2002). Between Chance and Choice, Interdisciplinary Perspectives on Determinism. Imprint Academic.Google Scholar
Avetisov, V. V. and Goldanskii, V. I. (1991). Homochirality and stereospecific activity: evolutionary aspects. Biosystems, 25 (3), 141–9.CrossRefGoogle ScholarPubMed
Ayala, F. J. (1983). Beyond Darwinism? The challenge of macroevolution to the synthetic theory of evolution. In PSA 1982: Proceedings of the 1982 Biennial Meeting of the Philosophy of Science Association Symposia, ed. Asquith, P. D. and Nickles, T., vol. 2, pp. 275–92.Google Scholar
Baas, N. A. (1994). Emergence, hierarchies, and hyperstructures. In Artificial Life III, Santa Fe Studies in the Science of Complexity, ed. Langton, C. G., vol. XVII Addison-Wesley, pp. 515–537.Google Scholar
Bachmann, P. A., Luisi, P. L., and Lang, J. (1992). Autocatalytic self-replication of micelles as models for prebiotic structures. Nature, 357, 57–9.CrossRefGoogle Scholar
Bachmann, P. A., Walde, P., Luisi, P. L., and Lang, J. (1990). Self-replicating reverse micelles and chemical autopoiesis. J. Am. Chem. Soc., 112, 8200–1.CrossRefGoogle Scholar
Bachmann, P. A., Walde, P., Luisi, P. L., and Lang, J. (1991). Self-replicating micelles: aqueous micelles and enzymatically driven reactions in reverse micelles. J. Am. Chem. Soc., 113, 8204–9.CrossRefGoogle Scholar
Bada, J. L. (1997). Meteoritics – extraterrestrial handedness?Science, 275, 942–3.CrossRefGoogle ScholarPubMed
Bada, J. F. and Lazcano, A. (2002). Some like it hot, but not the first biomolecules. Science, 296, 1982–3.CrossRefGoogle Scholar
Bada, J. L. and Lazcano, A. (2003). Prebiotic soup – revisiting the Miller experiment. Science, 300, 745–6.CrossRefGoogle ScholarPubMed
Baeza, I., Ibáñez, M., Santiago, J. C., et al. (1990). Diffusion of Mn2+ ions into liposomes mediated by phosphatidate and monitored by the activation of an encapsulated enzymatic system. J. Mol. Evol., 31, 453–61.CrossRefGoogle ScholarPubMed
Baeza, I., Wong, C., Mondragón, R., et al. (1994). Transbilayer diffusion of divalent cations into liposomes mediated by lipidic particles of phosphatidate, J. Mol. Evol., 39, 560–8.CrossRefGoogle ScholarPubMed
Bain, A. (1870). Logic, Books II and III. Longmans, Green & Co.Google Scholar
Bak, P., Tang, C., and Wisenfeld, K. (1988). Self-organized criticality. Physical Rev. A, 38, 364–74.CrossRefGoogle ScholarPubMed
Barbaric, S. and Luisi, P. L. (1981). Micellar solubilization of biopolymers in organic solvents. 5. Activity and conformation of α-chymorypsin in isooctane-AOT reverse micelles. J. Am. Chem. Soc., 103, 4239–44.CrossRefGoogle Scholar
Barrow, J. D. (2001). Cosmology, life and the anthropic principle. Ann. NY Acad. Sci., 950, 139–53.CrossRefGoogle ScholarPubMed
Barrow, J. D. and Tipler, F. J. (1986). The Anthropic Cosmological Principle. Oxford University Press.Google Scholar
Barrow, J. D. and Tipler, F. J. (1988). Action principles in nature. Nature, 331, 31–4.CrossRefGoogle Scholar
Bedau, M. A. (1997). Weak emergence. In Philosophical Perspectives: Mind, Causation and World, ed. Tomberlin, J.. Malden: Blackwell, vol. 11, pp. 375–99.Google Scholar
Jacob, Ben E., Becker, I., Shapira, Y., and Levine, H. (2004). Bacterial linguistic communication and social intelligence. Trends Microbiol., 12, 366–72.CrossRefGoogle ScholarPubMed
Benner, S. A. and Sismour, A. M. (2005). Synthetic biology. Nature Rev. Gen., 6, 524–45.CrossRefGoogle ScholarPubMed
Berclaz, N., Blöchliger, E., Müller, M., and Luisi, P. L. (2001a). Matrix effect of vesicle formation as investigated by cryotransmission electron microscopy. J. Phys. Chem. B, 105, 1065–71.CrossRefGoogle Scholar
Berclaz, N., Müller, M., Walde, P., and Luisi, P. L. (2001b). Growth and transformation of vesicles studied by ferritin labeling and cryotransmission electron microscopy. J. Phys. Chem. B, 105, 1056–64.CrossRefGoogle Scholar
Bernal, J. D. (1951). The Physical Basis of Life. Routledge & Paul.Google Scholar
Bernal, J. D. (1965), in Theoretical and Mathematical Biology, eds. Waterman, T. H. and Morowitz, H. J.. Blaisdell.Google Scholar
Bernal, J. D. (1967). The Origin of Life. World Publishing Company.Google Scholar
Bernal, J. D. (1971). Der Ursprung des Lebens. Editions Rencontre.Google Scholar
Bernard, C. (1865). Introduction to the Study of Experimental Medicine. Translated by H. C. Greene (1927). Henry Schuman.Google Scholar
Berti, D., Luisi, P. L., and Baglioni, P. (2000). Molecular recognition in supramolecular structures formed by phosphatidylnucleosides-based amphiphiles. Colloids Surf. A, 167, 95–103.CrossRefGoogle Scholar
Berti, D., Baglioni, P., Bonaccio, S., Barsacchi-Bo, G., and Luisi, P. L. (1998). Base complementarity and nucleoside recognition in phosphatidylnucleoside vesicles. J. Phys. Chem. B, 102, 303–8.CrossRefGoogle Scholar
Bianucci, M., Maestro, M., and Walde, P. (1990). Bell-shaped curves of the enzyme-activity in reverse micelles – a simplified model for hydrolytic reactions. Chem. Phys., 141, 273–83.CrossRefGoogle Scholar
Biebricher, K., Eigen, M., and Luce, R. (1981). Kinetic analysis of template, instructed and de novo RNA synthesis by Qbeta replicase. J. Mol. Biol., 148, 391–410.CrossRefGoogle Scholar
Billmeyer, F. W. (1984). Textbook of Polymer Science, 3rd edn. Wiley & Sons.Google Scholar
Birdi, K. S. (1999). Self-Assembly Monolayer Structures of Lipids and Macromolecules At Interfaces. Plenum Press.Google Scholar
Bissel, R. A., Cordova, E., Kaifer, A. E., and Stoddart, J. F. (1994). A chemically and electrochemically switchable molecular shuttle. Nature, 369, 133.CrossRefGoogle Scholar
Bitbol, M. (2001). Non-representationalist theories of knowledge and quantum mechanics. SATS, Nordic J. Phil., 2, 37–62.Google Scholar
Bitbol, M. and Luisi, P. L. (2004). Autopoiesis with or without cognition: defining life at its edge. J. Royal. Soc. Interface, 1, 99–107.CrossRefGoogle ScholarPubMed
Blocher, M., Hitz, T., and Luisi, P. L. (2001). Stereoselectivity in the oligomerization of racemic Tryptophan N-Carboxyanhydride (NCA-Trp) as determined by isotopic labelling and mass spectrometry. Helv. Chim. Acta, 84, 842–8.3.0.CO;2-1>CrossRefGoogle Scholar
Blocher, M., Liu, D., and Luisi, P. L. (2000). Liposome-assisted selective polycondensation of α-amino acids and peptides: the case of charged liposomes. Macromolecules, 33, 5787–96.CrossRefGoogle Scholar
Blocher, M., Walde, P., and Dunn, I. J. (1999). Modeling of enzymatic reactions in vesicles: the case of alpha-chymotrypsin. J. Biotechnol. Bioeng., 62, 36–43.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Blöchiger, E., Blocher, M., Walde, P., and Luisi, P. L. (1998). Matrix effect in the size distribution of fatty acid vesicles. J. Phys. Chem., 102, 10383–90.CrossRefGoogle Scholar
Böhler, C., Bannwarth, W., and Luisi, P. L. (1993). Self-replication of oligonucleotides in reverse micelles. Helv. Chim. Acta, 76, 2313–20.CrossRefGoogle Scholar
Böhringer, M., Morgenstern, K., Schneider, W. D., and Berndt, R. (1999). Separation of a racemic mixture of two-dimensional molecular clusters by scanning tunneling microscopy. Angew. Chem., Int. Ed. Engl., 38, 821–3.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Boiteau, L., Plasson, R., Collet, H., et al. (2002). Molecular origin of life: when chemistry became cyclic. The primary pump, a model for prebiotic emergence and evolution of petides. In Fundamentals of Life, eds. Palyi, G., Zucchi, C., and Caglioti, L.. Elsevier, pp. 211–18.Google Scholar
Boicelli, C. A., Conti, F., Giomini, M., and Giuliani, A. M. (1982). Interactions of small molecules with phospholipids in inverted micelles. Chem. Phys. Lett., 89, 490–6.CrossRefGoogle Scholar
Bolli, M., Micura, R., and Eschenmoser, A. (1997a). Pyranosyl-RNA: chiroselective self-assembly of base sequences by ligative oligomerization of tetranucleotide-2′,3′-cyclophosphates (with a commentary concerning the origin of biomolecular homochirality). Chem. Biol., 4, 309–20.CrossRefGoogle Scholar
Bolli, M., Micura, R., Pitsch, S., and Eschenmoser, A. (1997b). Pyranosyl-RNA: further observations on replication. Helv. Chim. Acta, 80, 1901–51.CrossRefGoogle Scholar
Bonaccio, S., Walde, P., and Luisi, P. L. (1994a). Liposomes containing purine and pyrimidine bases: stable unilamellar liposomes from phosphatidyl nucleosides. J. Phys. Chem., 98, 6661–3.CrossRefGoogle Scholar
Bonaccio, S., Cescato, C., Walde, P., and Luisi, P. L. (1994b). Self-production of supramolecular structures. In Liposomes from Lipidonucleotides and from Lipidopeptides, eds Fleischaker, G. R.et al.Kluwer Academic, pp. 225–59.Google Scholar
Bonaccio, S., Capitani, D., Segre, A. L., Walde, P., and Luisi, P. L. (1997). Liposomes from phosphatidyl nucleosides: an NMR investigation. Langmuir, 13, 1952–6.CrossRefGoogle Scholar
Bonaccio, S., Wessicken, M., Berti, D., Walde, P., and Luisi, P. L. (1996). Relation between the molecular structure of phosphatidyl nucleosides and the morphology of their supramolecular and mesoscopic aggregates. Langmuir, 12, 4976–78.CrossRefGoogle Scholar
Bonner, W. A. (1999). Chirality amplification – the accumulation principle revisited. Orig. Life Evol. Biosph., 29, 615–23.CrossRefGoogle ScholarPubMed
Böttcher, B., Lucken, U., and Graber, P. (1995). The structure of the H+-ATPase from chloroplasts by electron cryomicroscopy. Biochem. Soc. Trans., 23, 780–5.CrossRefGoogle ScholarPubMed
Bourgine, P. and Stewart, J. (2004). Autopoiesis and cognition. Artificial Life, 10 (3), 327–45.CrossRefGoogle ScholarPubMed
Brack, A. (ed.) (1998). The Molecular Origin of Life. Cambridge University Press.CrossRefGoogle Scholar
Brasier, M. D., Green, O. R., Jephcoat, A. P., et al. (2002). Questioning the evidence for Earth's oldest fossils. Nature, 416, 76–7.CrossRefGoogle ScholarPubMed
Briggs, T. and Rauscher, W. (1973). An oscillating iodine clock. J. Chem. Educ., 50, 496.CrossRefGoogle Scholar
Britt, R. R. (2000). Are we all aliens? The new case for panspermia. http://www.space.com.
Broad, C. D. (1925). The Mind and its Place in Nature. Routledge and Kegan.Google Scholar
Bucknall, D. G. and Anderson, H. L. (2003). Polymers get organized. Science, 302, 1904–5.CrossRefGoogle ScholarPubMed
Buhse, T., Lavabre, D., Nagarajan, R., and Micheau, J. C. (1998). Origin of autocatalysis in the biphasic alkaline hydrolysis of C-4 to C-8 ethyl alkanoates. J. Phys. Chem. A, 102, 10552–9.CrossRefGoogle Scholar
Buhse, T., Nagarajan, R., Lavabre, D., and Micheau, J. C. (1997). Phase-transfer model for the dynamics of “micellar autocatalysis”. J. Phys. Chem. A, 101, 3910–17.CrossRefGoogle Scholar
Bujdak, J., Eder, A., Yongyai, Y., Faybikova, K., and Rode, B. M. (1995). Peptide chain elongation: a possible role of montmorillonite in prebiotic synthesis of protein precursors. Orig. Life Evol. Biosph., 5, 431–41.CrossRefGoogle Scholar
Bujdak, J., Slosiarikova, H., Texler, N., Schwendinger, M., and Rode, B. M. (1994). On the possible role of montmorillonites in prebiotic peptide formation. Monats. Chem., 125, 1033–9.CrossRefGoogle Scholar
Burmeister, J. (1998). Self-replication and autocatalysis. In The Molecular Origin of Life, ed. Brack, A.. Cambridge University Press, pp. 295–310.CrossRefGoogle Scholar
Cairns-Smith, A. G. (1977). Takeover mechanisms and early biochemical evolution. Biosystems, 9, 105–9.CrossRefGoogle ScholarPubMed
Cairns-Smith, A. G. (1978). Precambrian solution photochemistry, inverse segregation, and banded iron formations. Nature, 276, 808–9.CrossRefGoogle Scholar
Cairns-Smith, A. G. (1982). Genetic Takeover and the Mineral Origins of Life. Cambridge University Press.Google Scholar
Cairns-Smith, A. G. (1990). Seven Clues to the Origin of Life. Cambridge University Press.Google Scholar
Cairns-Smith, A. G. and Walker, G. L. (1974). Primitive metabolism. Curr. Mod. Biol., 5 (4), 173–86.Google Scholar
Cairns-Smith, A. G., Hall, A. J., and Russell, M. J. (1992). Mineral theories of the origin of life and an iron sulphide example. Orig. Life Evol. Biosph., 22, 161–80.CrossRefGoogle Scholar
Calderone, C. T. and Liu, D. R. (2004). Nucleic acid-templated synthesis as a model system for ancient translation. Curr. Opin. Chem. Biol., 8, 645–53.CrossRefGoogle ScholarPubMed
Capra, F. (2002). The Hidden Connections. Harper Collins.Google Scholar
Carey, M. V. and Small, D. M. (1972). Micelle formation by bile salts. Physical-chemical and thermodynamic considerations. Arch. Intern. Med., 130, 506–27.CrossRefGoogle ScholarPubMed
Carr, B. (2001). Life, the cosmos and everything. Phys. World, 14, 23–5.CrossRefGoogle Scholar
Caselli, M., Maestro, M., and Morea, G. (1988). A simplified model for protein inclusion in reverse micelles. SANS measurements as a control test. Biotech. Prog., 4, 102–6.CrossRefGoogle Scholar
Cello, J., Paul, A. V., and Wimmer, E. (2002). Chemical synthesis of poliovirus cDNA: generation of infectious virus in the absence of natural template. Science, 297, 1016–18.CrossRefGoogle ScholarPubMed
Celovsky, V. and Bordusa, F. (2000). Protease-catalyzed fragment condensation via substrate mimetic strategy: a useful combination of solid-phase peptide synthesis with enzymatic methods. J. Pept. Res., 55, 325–9.CrossRefGoogle Scholar
Cevc, G. (1992). In Liposome Dermatics, eds. Braun-Falco, O., Korting, H. C., and Maibach, H. I.. Springer Verlag, pp. 82–90.CrossRefGoogle Scholar
Chakrabarti, A. C., Breaker, R. R., Joye, G. F., and Deamer, D. W. (1994). Production of RNA by a polymerase protein encapsulated within phospholipid vesicles. J. Mol. Evol., 39, 555–9.CrossRefGoogle ScholarPubMed
Chapman, K. B. and Szostak, J. W. (1995). Isolation of a ribozyme with 5'-5' ligase activity. Chem. Biol., 2, 325–33.CrossRefGoogle ScholarPubMed
Chen, I. A. and Szostack, J. W. (2004). A kinetic study of the growth of fatty acid vesicles. Bioph. J., 87, 988–98.CrossRefGoogle ScholarPubMed
Chen, I. A., Roberts, R. W., and Szostak, J. W. (2004). The Emergence of competition between model protocells. Science, 305, 1474–6.CrossRefGoogle ScholarPubMed
Cheng, Z. and Luisi, P. L. (2003). Coexistence and mutual competition of vesicles with different size distributions. J. Phys. Chem. B, 107 (39), 10940–5.CrossRefGoogle Scholar
Christidis, T. (2002). Probabilistic causality and irreversibility: Heraclitus and Prigogine. In Between Chance and Choice, eds. Atmanspacher, H. and Bishop, R.. Academic Imprint.Google Scholar
Chyba, C. F. and Sagan, C. (1992). Endogenous production, exogenous delivery and impact-shock synthesis of organic molelcules: an inventory for the origin of life. Nature, 355, 125–32.CrossRefGoogle Scholar
Chyba, F. and McDonald, G. D. (1995). The origin of life in the solar system: current issues. Ann. Rev. Earth Planet. Sci., 23, 215–49.CrossRefGoogle ScholarPubMed
Cistola, D. P., Hamilton, J. A., Jackson, D., and Small, D. M. (1988). Ionization and phase-behavior of fatty-acids in water. Application of the Gibbs phase rule. Biochemistry, 27, 1881–8.CrossRefGoogle ScholarPubMed
Commeyras, A., Boiteau, L., Vandenabeele-Trambouze, O., and Selsis, F. (2005). From prebiotic chemistry to the origins of life on Earth. In Lectures in Astrobiology, eds. Gargaud, M., Barbier, B., Martin, H. and Reisse, J.. Springer-Verlag, vol. I, part II, pp. 35–55.CrossRefGoogle Scholar
Conway-Morris, S. (2003). Life's Solution, Inevitable Humans in a Lonely Universe. Cambridge University Press.CrossRefGoogle Scholar
Cooper, G. W., Onwo, W. M., and Cronin, J. R. (1992). Alkyl phosphonic acids and sulfonic acids in the Murchison meteorite. Geochim. cosmochim. acta, 56, 4109–15.CrossRefGoogle ScholarPubMed
Corliss, J. B., Baross, J. A., and Hoffman, S. E. (1981). An hypothesis concerning the relationship between submarine hot springs and the origin of life. Oceanologica acta 4 Suppl., 59–69.Google Scholar
Corrin, M. L., Klevens, H. B., and Harkins, D. (1946). The determination of critical concentrations for the formation of soap micelles by the spectral behavior of pinacyanol chloride. J. Chem. Phys., 14, 480–6.CrossRefGoogle Scholar
Coveney, P. and Highfield, R. (1990). The Arrow of Time. W. H. Allen.Google Scholar
Crick, F. (1966). Of Molecules and Men. University of Washington Press.Google Scholar
Crick, F. (1980). The Astonishing Hypothesis. The Search of the Soul from a Chemical Perspective. Scribner.Google Scholar
Cronin, J. R. and Pizzarello, S. (1997). Enantiomeric excesses in meteoritic amino acids. Science, 275, 951–5.CrossRefGoogle ScholarPubMed
Crusats, J., Claret, J., Díez-Pérez, I., et al. (2003). Chiral shape and enantioselective growth of colloidal particles of self-assembled meso-tetra(phenyl and 4-sulfonatophenyl)porphyrins. Chem. Commun., 13, 1588–9.CrossRefGoogle Scholar
Cullis, P. R., Hope, M. J., Bally, M. B., et al. (1987). Liposomes as pharmaceuticals. In Liposomes. From Biophysics to Therapeutics, ed. Ostro, M. J.. Marcel Dekker, pp. 39–72.Google Scholar
Cullis, P. R., Hope, M. J., and Tilcock, C. P. S. (1986). Lipid polymorphism and the role of lipids in membranes. Chem. Phys. Lipids, 40, 127–44.CrossRefGoogle Scholar
Damasio, A. R. (1999). The Feeling of What Happens. Harcourt.Google Scholar
Davies, P. (1999). The Fifth Miracle: The Search for the Origin and Meaning of Life. Simon & Schuster.Google Scholar
Dawkins, R. (1990). The Blind Watchmaker: Why the Evidence of Evolution Reveals a Universe without Design. Penguin Books.Google Scholar
Dawkins, R. (2002). How Life Began: The Genesis of Life on Earth. Foundation for New Directions.Google Scholar
Deamer, D. W. (1985). Boundary structures are formed by organic components of the Murchison carbonaceous chondrite. Nature, 317, 792–4.CrossRefGoogle Scholar
Deamer, D. W. (1998). Possible starts for primitive life. In The Molecular Origins of Life, ed. Brack, A.. Cambridge University Press.Google Scholar
Deamer, D. W. and Pashley, R. M. (1989). Amphiphilic components of the Murchison carbonaceous chondrite: surface properties and membrane formation. Orig. Life Evol. Biosph., 19, 21–38.CrossRefGoogle ScholarPubMed
Deamer, D. W., Harang-Mahon, E., and Bosco, G. (1994). Self-assembly and function of primitive membrane structures. In Early Life on Earth. Nobel Symposium No. 84, ed. Bengston, S.. Columbia University Press, pp. 107–123.Google Scholar
Decher, G. (1997). Fuzzy nano-assemblies: toward layered polymeric multicomposites. Science, 277, 1232–7.CrossRefGoogle Scholar
Decker, P., Schweer, H., and Pohlmann, R. (1982). Identification of formose sugars, presumable prebiotic metabolites, using capillary gas chromatography/gas chromatography-mass spectrometry of n-butoxime trifluoroacetates on OV-225. J. Chromatogr., 225, 281–91.CrossRefGoogle Scholar
Duve, C. (1991). Blueprint for a Cell : The Nature and the Origin of Life. Neil Patterson Publishers.Google Scholar
Duve, C. (2002). Life Evolving: Molecules, Mind and Meaning. Oxford University Press.Google Scholar
Duve, C. and Miller, S. (1991). Two-dimensional life?Proc. Natl. Acad. Sci., 88, 10014–17.CrossRefGoogle ScholarPubMed
Feyter, S., Gesquiere, A., Wurst, K., et al. (2001). Homo- and heterochiral supramolecular tapes from achiral, enantiopure, and racemic promesogenic formamides: expression of molecular chirality in two and three dimensions. Angew. Chem. Int. Ed. Eng., 40, 3217–20.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Kruijff, B., Cullis, P. R. and Verkleij, A. J. (1980). Non-bilayer lipid structures in model and biological membranes. Trends Bioch. Sci., 5, 79–81.CrossRefGoogle Scholar
Napoli, M., Nardis, S., Paolesse, R., (2004). Hierarchical porphyrin self-assembly in aqueous solution. J. Am. Chem. Soc., 126, 5934–5.CrossRefGoogle ScholarPubMed
Ding, P. Z., Kawamura, K., and Ferris, J. P. (1996). Oligomerization of uridine phosphorimidazolides on montmorillonite: a model for the prebiotic synthesis of RNA on minerals. Orig. Life Evol. Biosph., 26, 151–71.CrossRefGoogle ScholarPubMed
Domazou, A. S. and Luisi, P. L. (2002). Size distribution of spontaneously formed liposomes by the alcohol injection method. J. Liposome Res., 12 (3), 205–20.CrossRefGoogle ScholarPubMed
Douglas, S., Zauner, S., Fraunholz, M., et al. (2001). The highly reduced genome of an enslaved algal nucleus. Nature, 410, 1091–2.CrossRefGoogle ScholarPubMed
Dubois, L. H. and Nuzzo, R. G. (1992). Synthesis, structure, and properties of model organic-surfaces. Ann. Rev. Phys. Chem., 43, 437–63.CrossRefGoogle Scholar
Dworkin, J. D., Deamer, D. W., Sandford, S., and Allmandola, L. (2001). Self-assembling amphiphilic molecules: synthesis in simulated interstellar/precometary ices. Proc. Natl. Acad. Sci., 98, 815–19.CrossRefGoogle ScholarPubMed
Dyson, F. J. (1982). A model for the origin of life. J. Mol. Evol., 18, 344–50.CrossRefGoogle ScholarPubMed
Dyson, F. J. (1985). Origins of Life. Cambridge University Press.Google Scholar
Ehrenfreund, P., Irvine, W., Becker, L., et al. (2002). Astrophysical and astrochemical insights into the origin of life. Rep. Prog. Phys., 65, 1427–87.CrossRefGoogle Scholar
Eichhorn, U., Bommarius, A. S., Drauz, K., and Jakubke, H.-D. (1997). Synthesis of dipeptides by suspension-to-suspension conversion via thermolysin catalysis: from analytical to preparative scale. J. Pept. Sci., 3, 245–51.3.0.CO;2-L>CrossRefGoogle ScholarPubMed
Eigen, M. (1971). Self-organization of matter and the evolution of biological macromolecules. Naturwissenschaften, 58, 465–523.CrossRefGoogle Scholar
Eigen, M. and Schuster, P. (1977). Hypercycle – principle of natural self-organization. A. Emergence of hypercycle. Naturwissenschaften, 64, 541–65.CrossRefGoogle ScholarPubMed
Eigen, M. and Schuster, P. (1979). The Hypercycle: a Principle of Natural Self-Organization. Springer Verlag.CrossRefGoogle Scholar
Eigen, M. and Winkler-Oswatitisch, R. (1992). Steps Towards Life. Oxford University Press.Google Scholar
El Seoud, O. A. (1984). In Reverse Micelles, eds. Luisi, P. L. and Straub, B.. Plenum Press.CrossRefGoogle Scholar
Elias, H. G. (1997). An Introduction to Polymer Science. Wiley VCH.Google Scholar
Engels, F. (1894). In Herrn Eugen Dühring's Umwalzung der Wissenschaft, Dietz Verlag.Google Scholar
Epstein, S., Krishnamurthy, R. V., Cronin, J. R., Pizzarello, S., and Yuen, G. U. (1987). Unusual stable isotope ratios in amino acid and carboxylic-acid extracts from the Murchison meteorite. Nature, 326, 477–9.CrossRefGoogle ScholarPubMed
Erickson, J. C. and , Kennedy R. M. (1980). Effects of histidyl-histidine and polyribonucleotides on glycine condensation in fluctuating clay environments. Abstracts Papers Am. Chem. Soc., 179, 43.Google Scholar
Ericsson, B., Larsson, K., and Fontell, K. (1983). A cubic protein-monoolein-water phase. Biochim. Biophys. Acta, 729, 23–7.CrossRefGoogle ScholarPubMed
Erwin, D. H. (2003). Life's solution – inevitable humans in a lonely universe. Science, 302, 1682–3.CrossRefGoogle Scholar
Eschenmoser, A. (1999). Chemical etiology of nucleic acid structure. Science, 284, 2118–24.CrossRefGoogle ScholarPubMed
Eschenmoser, A. (2003). Creating a perspective for comparing. In Proceedings of the J. Templeton Foundation “Biochemistry and Fine-tuning”, Harvard University, October 10–12, 2003.Google Scholar
Eschenmoser, A. and Kisakürek, M. V. (1996). Chemistry and the origin of life. Helv. Chim. Acta, 79, 1249–59.CrossRefGoogle Scholar
Fadnavis, N. W. and Luisi, P. L. (1989). Immobilized enzymes in reverse micelles: studies with gel-entrapped Trypsin and alpha-Chymotrypsin in AOT reverse micelles. Biotechnol. Bioeng., 33, 1277–82.CrossRefGoogle ScholarPubMed
Falbe, J. (1987). Surfactants in Consumer Products. Theory, Technology and Applications. Springer Verlag.CrossRefGoogle Scholar
Famiglietti, M., Hochköppler, A., and Luisi, P. L. (1993). Surfactant-induced hydrogen production in cyanobacteria. Biotechnol. Bioeng., 42, 1014–18.CrossRefGoogle ScholarPubMed
Famiglietti, M., Hochköppler, A., Wehrli, E., and Luisi, P. L. (1992). Photosynthetic activity of cyanobacteria in water-in-oil microemulsions. Biotechnol. Bioeng., 40, 173–8.CrossRefGoogle ScholarPubMed
Farre, L. and Oksala, T., eds. (1998). Emergency, complexity, hierarchy, organisation. Selected papers from the ECHO III Conference (ESPOO, Finland), Acta Polytechnica Scandi., 91.Google Scholar
, Fendler J. H. and Fendler, E. J. (1975). Catalysis in Micellar and Macromolecular Systems. Academic Press.Google Scholar
Ferris, J. P. (1998). Catalyzed RNA synthesis for the RNA world. In The Molecular Origin of Life, ed. Brack, A.. Cambridge University Press, pp. 255–6.CrossRefGoogle Scholar
Ferris, J. P. (2002). From building blocks to the polymers of life. In Life's Origin, The Beginning of Biological Evolution, ed. Schopf, J. W.. California University Press, 113–39.Google Scholar
Ferris, J. P. and Ertem, G. (1992). Oligomerization reaction of ribonucleosides on montmorillonite: reaction of 5′-phosphorimidazolide of adenosine. Science, 257, 1387–9.CrossRefGoogle Scholar
Ferris, J. P. and Ertem, G. (1993). Montmorillonite catalysis of RNA oligomer formation in aqueous solution: a model for the prebiotic formation of RNA. J. Am. Chem. Soc., 115, 12270–5.CrossRefGoogle ScholarPubMed
Ferris, J. P., Donner, D. B., and Lobo, A. P. (1973). Possible role of hydrogen cyanide in chemical evolution. The oligomerization and condensation of hydrogen cyanide. J. Mol. Biol., 74, 511–18.CrossRefGoogle ScholarPubMed
Ferris, J. P., Sanchez, R. A., and Orgel, L. E. (1968). Studies in prebiotic synthesis. III, Synthesis of pyrimidines from cyanoacetilene and cyanate. J. Mol. Biol., 33, 693–704.CrossRefGoogle Scholar
Ferris, J. P., Hill, R. Jr., Liu, R., and Orgel, L. (1996). Synthesis of long prebiotic oligomers on mineral surface. Nature, 381, 59–61.CrossRefGoogle Scholar
Ferris, J. P., Joshi, P. C., Edelson, E. H., and Lawless, J. G. (1978). HCN: a plausible source of purines, pyrimidines and amino acids on the primitive earth. J. Mol. Evol., 11, 293–311.CrossRefGoogle ScholarPubMed
Ferris, J. P., Wos, J. D., Nooner, D. W., and Oró, J. (1974). Chemical Evolution. 21. Amino-Acids Released on Hydrolysis of HCN Oligomers. J. Mol. Evol., 3, 225–31.CrossRefGoogle Scholar
Field, R. J. (1972). A reaction periodic in time and space. J. Chem. Educ., 49, 308–11.CrossRefGoogle Scholar
Fischer, A., Franco, A., and Oberholzer, T. (2002). Giant vesicles as microreactors for enzymatic mRNA synthesis. Chem. Bio. Chem., 3 (5), 409–17.3.0.CO;2-P>CrossRefGoogle ScholarPubMed
Fischer, A., Oberholzer, T., and Luisi, P. L. (2000). Giant vesicles as models to study the interactions between membranes and proteins. Biochim. Biophys. Acta, 1467, 177–88.CrossRefGoogle Scholar
Fleischaker, G. (1988). Autopoiesis: the status of its system logic. Biosystems, 22, 37–49.CrossRefGoogle ScholarPubMed
Fletcher, P. D. and Robinson, B. H. (1981). Ber. Bunsenges. Phys. Chem., 85, 863.CrossRef
Foldvari, M., Geszles, A., and Mezei, M. (1990). J. Microencapsul., 7, 479–89.CrossRef
Folsome, C. E. (1979). The Origin of Life: A Warm Little Pond. W. H. Freeman & Co.Google Scholar
Fontell, K. (1990). Cubic phases in surfactant and surfactant-like lipid systems. Colloid Polym. Sci., 268, 265–85.CrossRefGoogle Scholar
Föster, S. and Plantenberg, T. (2002). From self-organizing polymers to nanohybrid and biomaterials. Angew. Chem. Int. Ed. Engl., 41, 688–714.3.0.CO;2-3>CrossRefGoogle Scholar
Fox, S. W. (1988). The Emergence of Life. Basic Books.Google Scholar
Fox, S. W. and Dose, K. (1972). Molecular Evolution and the Origin of Life. W. H. Freeman.Google Scholar
Fox, S. W. and Dose, K. (1977). Molecular Evolution and the Origin of Life. New York and Basel: Marcel Dekker.Google Scholar
Franchi, M. and Gallori, E. (2004). Origin, persistence and biological activity of genetic material in prebiotic habitats. Orig. Life Evol. Biosph., 34, 133–41.CrossRefGoogle ScholarPubMed
Franchi, M., Ferris, J. P., and Gallori, E. (2003). Cations as mediators of the adsorption of nucleic acids on clay surfaces in prebiotic environments. Orig. Life Evol. Biosph., 33, 1–16.CrossRefGoogle ScholarPubMed
Franz, M.-L. (1988). Psyche und Materie. Daimon Verlag.Google Scholar
Fraser, C. M., Gocayne, J. D., White, O., et al. (1995). The minimal gene complement of Mycoplasma genitalium. Science, 270, 397–403.CrossRefGoogle ScholarPubMed
Freitas, R. A. Jr. and Merkle, R. C. (2004). Kinematic Self-Replicating Machines. Landes Bioscience.Google Scholar
Frick, D. N. and Richardson, C. C. (2001). DNA Primases. Annu. Rev. Biochem., 70, 39–80.CrossRefGoogle ScholarPubMed
Funqua, C., Parsek, M. R., and Greenberg, E. P. (2001). Regulation of gene expression by cell-to-cell communication: acyl-homoserine lactone quorum sensing. Ann. Rev., Genet., 35, 439–68.Google Scholar
Ganti, T. (1975). Organization of chemical reactions into dividing and metabolizing units: the chemotons. BioSystems, 7, 15–21.CrossRefGoogle ScholarPubMed
Ganti, T. (1984). Chemoton elmélet 1. kötet. A fluid automaták elméleti alapjai (Translated as Chemoton Theory. Vol. 1. Theory of Fluid Automata). OMIKK.Google Scholar
Ganti, T. (2003). The Principles of Life. Oxford University Press.CrossRefGoogle Scholar
Gao, X. and Huang, L. (1995). Cationic liposome-mediated gene transfer, Gene Ther., 2 (10), 710–22.Google ScholarPubMed
Gavrilova, L. P., Kostiashkina, O. E., Koteliansky, V. E., Rutkevitch, N. M., and Spirin, A. S. (1976). Factor-free (non-enzymic) and factor-dependent systems of translation of polyuridylic acid by E. coli ribosomes. J. Mol. Biol., 101, 537–52.CrossRefGoogle Scholar
Gennis, R. B. (1989). Biomembranes, Molecular Structure and Function. Springer Verlag.Google Scholar
Ghosh, I. and Chmielewski, J. (2004). Peptide self-assembly as a model of proteins in the pre-genomic world. Curr. Opin. Chem. Biol., 8, 640–4.CrossRefGoogle ScholarPubMed
Gil, R., Silva, F. J., Peretó, J., and Moya, A. (2004). Determination of the core of a minimal bacteria gene set. Microb. Molec. Biol. Rev., 68, 518–37.CrossRefGoogle ScholarPubMed
Gil, R., Sabater-Munoz, B., Latorre, A., Silva, F. J., and Moya, A. (2002). Extreme genome reduction in Buchnera spp: toward the minimal genome needed for symbiotic life. Proc. Natl. Acad. Sci. USA, 99, 4454–8.CrossRefGoogle ScholarPubMed
Gilbert, W. (1986). The RNA world. Nature, 319, 618.CrossRefGoogle Scholar
Glade, N., Demongeot, J., and Tabony, J. (2004). Microtubules self-organization by reaction-diffusion processes causes collective transport and organization of cellular particles. BMC Cell Biol., 5, 5–23.CrossRefGoogle Scholar
Glotzer, S. C. (2004). Materials science. Some assembly required. Science, 306, 419–20.CrossRefGoogle ScholarPubMed
Gould, S. J. (1989). Wonderful Life. Penguin Books.Google Scholar
Graf, A., Winterhalter, M., and Meier, W. (2001). Nanoreactors from polymer-stabilized liposomes. Langmuir, 17, 919–23.CrossRefGoogle Scholar
Gregoriadis, G. (ed.) (1988). Liposomes ad Carriers of Drugs: Recent Trends and Progress. Wiley.Google Scholar
Gregoriadis, G. (1976a). The carrier potential of liposomes in biology and medicine (first of two parts). New Engl. J. Med., 295, 704–10.CrossRefGoogle Scholar
Gregoriadis, G. (1976b). The carrier potential of liposomes in biology and medicine (second of two parts). New Engl. J. Med., 295, 765–70.CrossRefGoogle Scholar
Gregoriadis, G. (1995). Engineering liposomes for drug delivery: progress and problems. Trends Biotechnol., 13 (12), 527–37.CrossRefGoogle ScholarPubMed
Häckel, E. (1866). Allgemeine Anatomie der Organismen. Walter de Gruyer.Google Scholar
Häring, G., Luisi, P. L., and , Hauser H. (1988). Characterization by electron spin resonance of reversed micelles consisting of the ternary system AOT–isooctan–water. J. Phys. Chem., 92, 3574–81.CrossRefGoogle Scholar
Häring, G., Luisi, P. L., and Meussdoerffer, F. (1985). Solubilization of bacteria cells in organic solvents via reverse micelles. Biochem. Biophys. Res. Commun., 127, 911–15.CrossRefGoogle ScholarPubMed
Häring, G., Pessina, A., Meussdoerffer, F., Hochkoppler, A., and Luisi, P. L. (1987). Solubilization of bacterial cells in organic solvents via reverse micelles and microemulsions. Ann. Biochem. Eng., 506, 337–344.Google Scholar
Haldane, J. B. S. (1929). The origin of life. Rationalist Annual, 148, 3–10.Google Scholar
Haldane, J. B. S. (1954). The origin of life. New Biol., 16, 12–27.Google Scholar
Haldane, J. B. S. (1967). In The Origin of Life, ed. Bernal, J. D.. World Publishing Co.Google Scholar
Halling, P. J., Eichhorn, U., Kuhl, P., and Jakubke, H.-D. (1995). Thermodynamics of solid-to-solid conversion and application to enzymic peptide synthesis. Enzyme Microb. Technol., 17, 601–6.CrossRefGoogle Scholar
Han, D. and Rhee, J. S. (1986). Biotechnol. Bioeng., 27, 1250–5.CrossRef
Hanczyc, M. M., Fujikawa, S. M., and Szostak, J. W. (2003). Experimental models of primitive cellular compartments: encapsulation, growth, and division. Science, 302, 618–22.CrossRefGoogle ScholarPubMed
Hansler, M. and Jakubke, H.-D. (1996). Nonconventional protease catalysis in frozen aqueous solutions. J. Pept. Sci., 2, 279–89.CrossRefGoogle ScholarPubMed
Harada, S. and Schelly, Z. A. (1982). Reversed micelle of dodecylpyridinium iodide in benzene. Pressure-jump relaxation kinetic and equilibrium study of the solubilization of 7,7,8,8-tetracyanoquinodimethane. J. Phys. Chem., 86, 2098–102.CrossRefGoogle Scholar
Hargreaves, W. R and Deamer, D. W. (1978a). Liposomes from ionic, single-chain amphiphiles. Biochemistry, 17, 3759–68.CrossRefGoogle Scholar
Hargreaves, W. R and Deamer, D. W. (1978b). In Light Transducing membranes: Structure, Function and Evolution, ed. Deamer, D. W.. Academic Press, pp. 23–59.Google Scholar
Hargreaves, W. R., Mulvhill, S. J., and Deamer, D. W. (1977). Synthesis of phospholipids and membranes in prebiotic conditions. Nature, 266, 78–80.CrossRefGoogle ScholarPubMed
Havinga, E. (1954). Spontaneous formation of optically active substances. Biochem. Biophys. Acta, 13, 171–4.CrossRefGoogle ScholarPubMed
Hawker, C. J. and Frechet, J. M. J. (1990). Preparation of polymers with controlled molecular architecture – a new convergent approach to dendritic macromolecules. J. Am. Chem. Soc., 112, 7638–47.CrossRefGoogle Scholar
Hayatsu, R., Studier, M. H., Moore, L. P., and Anders, E. (1975). Purines and triazines in the Murchison meteorite. Geochim. Cosmochim. Acta, 39, 471–88.CrossRefGoogle Scholar
Heinen, W. and Lauwers, A. M. (1997). The iron-sulfur world and the origins of life: abiotic thiol synthesis from metallic iron, H2S and CO2; a comparison of the thiol generating FeS/HCl(H2S)/CO2-system and its Fe0/H2S/CO2-counterpart. Proc. Royal Netherlands Acad. Arts Sci., 100, 11–25.Google Scholar
Hilborn, R. C. (1994). Chaos and Non Linear Dynamics. Oxford University Press.Google Scholar
Hilhorst, R., Spruijt, R., Laane, C., and Veeger, C. (1984). Rules for the regulation of enzyme-activity in reversed micelles as illustrated by the conversion of apolar steroids by 20-beta-hydroxysteroid dehydrogenase. Eur. J. Biochem., 144, 459–66.CrossRefGoogle ScholarPubMed
Hitz, T. and Luisi, P. L. (2004). Spontaneous onset of homochirality in oligopeptide chains generated in the polymerization of N-carboxyanhydride amino acids in water. Orig. Life Evol. Biosph., 34 (1–2), 93–110.CrossRefGoogle ScholarPubMed
Hochköppler, A. and Luisi, P. L. (1989). Solubilization of soybean mitochondria in AOT/isooctane water-in-oil microemulsions. Biotechnol. Bioeng., 33, 1477–81.CrossRefGoogle Scholar
Hochköppler, A. and Luisi, P. L. (1991). Photosynthetic activity of plant cells solubilized in water-in-oil microemulsions. Biotechnol. Bioeng., 37, 918–21.CrossRefGoogle Scholar
Hochköppler, A., Pfammatter, N., and Luisi, P. L. (1989). Activity of yeast cells solubilized in water-in-oil microemulsions. Chimia, 43, 348–350.Google Scholar
Holden, C. (2005). Vatican astronomer rebuts cardinals' attack on Darwinism. Science, 309, 996–7.CrossRefGoogle ScholarPubMed
Holland, J. H. (1998). Emergence: From Chaos to Order. Oxford University Press.Google Scholar
Holm, N. G. and Andersson, E. M. (1998). Hydrothermal systems. In The Molecular Origin of Life, ed. Brack, A.. Cambridge University Press.CrossRefGoogle Scholar
Holm, N. G., Cairns-Smith, A. G., Daniel, R. M., et al. (1992). Marine hydrothermal systems and the origin of life: future research. Orig. Life Evol. Biosph., 22, 181–242.CrossRefGoogle ScholarPubMed
Horowitz, N. and Miller, S. (1962). In Progress in the Chemistry of Natural Products, ed. Zechmeister, L.. Springer Verlag, vol. 20, pp. 423–59.Google Scholar
Horowitz, P. and Sagan, C. (1993). Five years of Project META: an all-sky narrow-band radio search for extraterrestrial signals. Astrophys. J., 415, 218–33.CrossRefGoogle Scholar
Howard, F. B., Frazier, J., Singer, M. F., and Miles, H. T. (1966). Helix formation between polyribonucleotides and purine nucleosides and nucleotides. 2. J. Mol. Biol., 16, 415.CrossRefGoogle Scholar
Hoyle, F. and Wickramasinghe, C. (1999). Astronomical origins of life – steps toward panspermia. Astrophys. Space Sci., 268, preface vii–viii.Google Scholar
Hoyle, F., and Wickramasinghe, C. (2000). Astronomical Origins of Life – Steps Towards Panspermia. Kluwer Academic.CrossRefGoogle Scholar
Huang, S. S. (1959). Occurrence of life in the universe. Amer. Sci., 47, 397–402.Google Scholar
Huang, W. M. and Tso, P. O. P. (1966). Physicochemical basis of recognition process in nucleic acid interactions. I. Interactions of polyuridylic acid and nucleosides. J. Mol. Biol., 16, 523.CrossRefGoogle ScholarPubMed
Huber, C. and Wächtershäuser, G. (1997). Activated acetic acid by carbon fixation on (Fe, Ni)S under primordial condition. Science, 276, 245–7.CrossRefGoogle Scholar
Hutchinson, C. A., Peterson, S. N., Gill, S. R., et al. (1999). Global transposon mutagenesis and a minimal Mycoplasma genome. Science, 286, 2165–9.CrossRefGoogle Scholar
Imre, V. E. and Luisi, P. L. (1982). Solubilization and condensed packaging of nucleic acids in reversed micelles. Biochem. Biophys. Res. Commun., 107, 538–45.CrossRefGoogle ScholarPubMed
Inoue, T. and Orgel, L. (1983). A nonenzymatic RNA polymerase model. Science, 219, 859–62.CrossRefGoogle ScholarPubMed
Ishikawa, K., Sato, K., Shima, Y., Urabe, I., and Yomo, T. (2004). Expression of cascading genetic network within liposomes. FEBS Lett., 576, 387–90.CrossRefGoogle ScholarPubMed
Islas, S., Becerra, A., Luisi, P. L., and Lazcano, A. (2004). Comparative genomics and the gene complement of a minimal cell. Orig. Life Evol. Biosph., 34 (1–2), 243–56.CrossRefGoogle ScholarPubMed
Israelachvili, J. N, Mitchell, D. J., and Ninham, B. W. (1977). Theory of self-assembly of lipid bilayers and vesicles. Biochim. Biophys. Acta, 470, 185–201.CrossRefGoogle ScholarPubMed
Israelachvili, J. N. (1992). Intermolecular and Surface Forces, 2nd edn. Academic Press.Google Scholar
Issac, R. and Chmielewski, J. (2002). Approaching exponential growth with a self-replicating peptide. J. Am. Chem. Soc., 124, 6808–9.CrossRefGoogle ScholarPubMed
Itaya, M. (1995). An estimation of the minimal genome size required for life. FEBS Lett., 362, 257–60.CrossRefGoogle ScholarPubMed
Ito, Y., Fujii, H., and Imanishi, Y. (1993). Catalytic peptide synthesis by trypsin modified with polystyrene in chloroform. Biotechnol. Prog., 9 (2), 128–30.CrossRefGoogle ScholarPubMed
Itojima, Y., Ogawa, Y., Tsuno, K., Handa, N., and Yanagawa, H. (1992). Spontaneous formation of helical structures from phospholipid-nucleoside conjugates. Biochemistry, 31, 4757–65.CrossRefGoogle ScholarPubMed
Jacob, F. (1982). The Possible and the Actual. University of Washington Press.Google Scholar
Jager, L., Wright, M. C., and Joyce, G. F. (1999). A complex ligase ribozyme evolved in vitro from a group I ribozymes domain, Proc. Natl. Acad. Sci., 96, 14712–17.CrossRefGoogle Scholar
Jakubke, H.-D. (1987). In The Peptides: Analysis, Synthesis, Biology, eds. Udenfried, S. and Meienhofer, J.. Academic Press, vol. 9, ch. 3.Google Scholar
Jakubke, H.-D. (1995). In Enzyme Catalysis in Organic Synthesis, eds. Drauz, K. and Waldmann, H.. VCH vol. 1, pp. 431–58.CrossRefGoogle Scholar
Jakubke, H.-D., Kuhl, P., and Könnecke, A. (1985). Basic principles of protease-catalyzed peptide bond formation. Angew. Chem. Int. Ed. Engl., 24, 85–93.CrossRefGoogle Scholar
Jakubke, H.-D., Eichhorn, U., Hansler, M., and Ullmann, D. (1996). Non-conventional enzyme catalysis: application of proteases and zymogens in biotransformations. Biol. Chem., 377, 455–64.Google ScholarPubMed
Janiak, M. J., Small, D. M., and Shipley, G. G. (1976). Nature of the thermal pretransition of synthetic phospholipids: dimyristoyl- and dipalmitoyllecithin. Biochemistry, 15, 4575–80.CrossRefGoogle ScholarPubMed
Jay, D. and Gilbert, W. (1987). Basic protein enhances the encapsulation of DNA into lipid vesicles: model for the formation of primordial cells. Proc. Natl. Acad. Sci. USA, 84, 1978–80.CrossRefGoogle Scholar
Jimenez-Prieto, R., Silva, M., and Perez-Bendito, D. (1998). Approaching the use of oscillating reactions for analytical monitoring. Analyst, 123, 1R–8R.CrossRefGoogle Scholar
Joyce, G. F. (1989). RNA evolution and the origin of life. Nature, 338, 217–24.CrossRefGoogle Scholar
Joyce, G. F. and Orgel, L. E. (1986). Nonenzymatic template-directed synthesis on RNA random copolymers – poly(C, G) templates. J. Mol. Biol., 188, 433–41.CrossRefGoogle Scholar
Joyce, G. F. and Orgel, L. E. (1993). Prospects for understanding the origin of the RNA World. In The RNA World, eds. Gesteland, R. F. and Atkins, J. F.. Plainview, Cold Spring Harbor Laboratory Press, pp. 1–25.Google Scholar
Joyce, J. (1994). In Origins of Life: The Central Concepts, eds. Deamer, D. W. and Fleischaker, G. R.. Jones and Bartlett, foreword.Google Scholar
Kaler, E. W., Murthy, A. K., Rodriguez, B. E., and Zasadzinski, J. A. N. (1989). Spontaneous vesicle formation in aqueous mixtures of single-tailed surfactants. Science, 245, 1371–4.CrossRefGoogle ScholarPubMed
Karnup, A. S., Uverskii, V. N., and Medvedkin, V. N. (1996). Synthetic Polyaminoacids and Polypeptides, Preparation by the N-carboxyanhydride method. Russ. J. Bioorg. Chem., 22, 479–90.Google Scholar
Kaszuba, M. and Jones, M. N. (1999). Hydrogen peroxide production from reactive liposomes encapsulating enzymes. Biochim. Biophys. Acta, 1419, 221–8.CrossRefGoogle ScholarPubMed
Kauffman, S. A. (1986). Autocatalytic set of proteins. J. Theor. Biol., 119, 1–24.CrossRefGoogle ScholarPubMed
Kauffman, S. A. (1993). The Origins of Order. Oxford: Oxford University Press.Google Scholar
Kawamura, K. and Kamoto, F. (2000). Condensation reaction of hexanucleotides containing guanine and cytosine with water soluble carbodiimide. Nucleic Acid Symp. Ser., 44, 217–18.CrossRefGoogle Scholar
Kawamura, K. (2002). The origin of life from the life of subjectivity. In Fundamentals of life, eds. Palyi, G., Zucchi, C., and Caglioti, L.. Elsevier, pp. 563–76.Google Scholar
Kent, S. (1999). Chemical protein synthesis by solid phase ligation of unprotected peptide segments. J. Am. Chem. Soc., 121, 8720–7.Google Scholar
Kessaissia, S., Siffert, B., and Donnet, J. B. (1980). Synthese de peptides; preparation de l'acide hippurique par reaction des complexes montmorillonite-glycine avec l'acide benzoique. Clay Minerals, 15, 383–92.CrossRefGoogle Scholar
Kiedrowski, G. (1986). A self-replicating hexadeoxynucleotide. Angew. Chem. Int. Ed. Engl., 25, 932–5.CrossRefGoogle Scholar
Kiedrowski, G. (1993). Minimal replicator theory I: parabolic versus exponential growth. In Bioorganic Chemistry, ed. Berlin, D. H.. Springer Verlag, vol. 3, pp. 115–46.Google Scholar
Kikuchi, A., Aoki, Y., Sugaya, S., et al. (1999). Development of novel cationic liposomes for efficient gene transfer into peritoneal disseminated tumor. Human. Gene Ther. 10 (6), 947–55.CrossRefGoogle ScholarPubMed
Kim, J. (1984). Concepts of supervenience. Phil. Phen. Res., 45, 153–76.CrossRefGoogle Scholar
Kimura, M. (1983). The Neutral Theory of Molecular Evolution. Cambridge University Press.CrossRefGoogle Scholar
Kirstein, S., Berlepsch, H., Bottecher, C., et al. (2000). Chiral J-aggregates formed by achiral cyanine dyes. Chem. Phys. Chem., 1, 146–50.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Klee, R. (1984). Micro-determinism and concepts of emergence. Phil. Sci., 51, 44–63.CrossRefGoogle Scholar
Knenvolden, K., Lawless, J. G., Pering, K., et al. (1970). Evidence for extraterrestrial amino acids and hydrocarbons in the Murchison meteorite. Nature, 228, 923–6.CrossRefGoogle Scholar
Kolisnychenko, V., Plunkett, G., Herring, C. D., et al. (2002). Engineering a reduced Escherichia coli genome. Genome Res., 12, 640–7.CrossRefGoogle ScholarPubMed
Kondepudi, D. K. and Prigogine, I. (1981). Sensitivity of non-equilibrium systems. Physica A, 107, 1–24.CrossRefGoogle Scholar
Kondepudi, D. K., Kaufman, R., and Singh, N. (1990). Chiral symmetry breaking in sodium chlorate crystallization. Science, 250, 975.CrossRefGoogle Scholar
Kondepudi, D. K., Prigogine, I., and Nelson, G. (1985). Sensitivity of branch selection in nonequilibrium systems. Phys. Lett. A, 111, 29–32.CrossRefGoogle Scholar
Kondo, Y., Uchiyama, H., Yoshino, N., Nishiyama, K., and Abe, M. (1995). Spontaneous vesicle formation from aqueous-solutions of didodecyldimethylammonium bromide and sodium dodecyl-sulfate mixtures. Langmuir, 11, 2380–4.CrossRefGoogle Scholar
Koonin, E. V. (2000). How many genes can make a cell: the minimal-gene-set concept. Annu. Rev. Genomics Human Genet. 1, 99–116.CrossRefGoogle ScholarPubMed
Koshland, D. E. Jr. (2002). The seven pillars of life. Science. 295, 2215–16.CrossRefGoogle ScholarPubMed
Koster, G., Duijin, M., Hofs, B., and Dogterom, M. (2003). Membrane tube formation from giant vesicles by dynamic association of motor proteins. Proc. Natl. Acad. Sci., 100, 15583–8.CrossRefGoogle ScholarPubMed
Kricheldorf, H. R. (1990). In Models of Biopolymers by Ring-Opening Polymerization, ed. Penczek, S.. CRC Press, pp. 46–62.Google Scholar
Kricheldorf, H. R. and Hull, W. E. (1979). Stereospecificity of the polymerization of D, L-alanine-NCA and D, L-alanine NCA. Makromol. Chem., 180, 1715–24.CrossRefGoogle Scholar
Kuiper, T. B. H. and Morris, M. (1977). Searching for extraterrestial civilizations. Science, 196, 616–21.CrossRefGoogle Scholar
Kullmann, W. (1987). Enzymatic Peptide Synthesis. CRC Press.Google Scholar
Lahav, M. and Leiserowitz, L. (1999). Spontaneous resolution: from three-dimensional crystals to two-dimensional magic nanoclusters. Angew. Chem. Int. Ed. Engl., 38, 2533–6.3.0.CO;2-B>CrossRefGoogle ScholarPubMed
Landau, E. M. and Luisi, P. L. (1993). Lipid cubic phases as transparent, rigid matrices for the direct spectroscopic study of immobilized membrane proteins. J. Am. Chem. Soc., 115, 2102–6.CrossRefGoogle Scholar
Landweber, L. F. and Pokrovskaya, I. D. (1999). Emergence of a dual-catalytic RNA with metal-specific cleavage and ligase activities: the spandrels of RNA evolution. Proc. Natl. Acad. Sci., 96, 173–8.CrossRefGoogle ScholarPubMed
Langton, C. G. (1990). Computation at the edge of chaos: phase transitions and emergent computation. Physica, D42, 12–37.Google Scholar
Larsson, K. (1989). Cubic lipid-water phases: structures and biomembrane aspects. J. Phys. Chem., 93, 7304–14.CrossRefGoogle Scholar
Lasch, J., Laub, R., and Wohlrab, W. (1991). How deep do intact liposomes penetrate into human skin?J. Controll. Release, 18, 55–8.CrossRefGoogle Scholar
Lasic, D. D. (1995). In Handbook of Biological Physics, eds. Lipowsky, R. and Sackmann, E.. Elsevier vol. 1, pp. 491–519.Google Scholar
Lauceri, R., Raudino, A., Scolaro, L. M., Mical, N., and Purrello, R. (2002). From achiral porphyrins to template-imprinted chiral aggregates and further: self-replication of chiral memory from scratch. J. Am. Chem. Soc., 124, 894–5.CrossRefGoogle ScholarPubMed
Lawless, J. G. and Yuen, G. U. (1979). Quantification of monocarboxylic acids in the Murchison carbonaceous meteorite. Nature, 282, 396–8.CrossRefGoogle Scholar
Lawrence, D. S., Jiang, T., and Levett, M. (1995). Self-assembling supramolecular complexes. Chem. Rev., 95, 2229–60.CrossRefGoogle Scholar
Lazcano, A. (2003). Just how pregnant is the universe?Science, 299, 347–8.CrossRefGoogle Scholar
Lazcano, A. (2004). An answer in search of a question how life began: the genesis of life on Earth, by William Day. Astrobiology, 4, 469–71.CrossRefGoogle Scholar
Lazcano, A. (2005). Teaching evolution in Mexico: preaching to the choir. Science, 310, 787–9.CrossRefGoogle ScholarPubMed
Lazcano, A. and Bada, J. L. (2003). The 1953 Stanley L. Miller experiment: fifty years of prebiotic organic chemistry. Orig. Life Evol. Biosph., 33, 235–42.CrossRefGoogle Scholar
Lazcano, A., Guerriero, R., Margulius, L., and Oró, J. (1988). The evolutionary transition from RNA to DNA in early cells. J. Mol. Evol., 27, 283–90.CrossRefGoogle ScholarPubMed
Lazcano, A., Valverde, V., Hernandez, G., et al. (1992). On the early emergence of reverse transcription: theoretical basis and experimental evidence. J. Mol. Evol., 35, 524–36.CrossRefGoogle ScholarPubMed
Lazzara, S. (2001). Vedi Alla Voce Scienza. Manifesto Libri.Google Scholar
Doux, J. (2002). Synaptic Self : How Our Brains Become Who We Are. Viking Books.Google Scholar
Lee, D. H., Granja, J. R., Martinez, J. A., Severin, K., and Ghadiri, M. R. (1996). A self-replicating peptide. Nature, 382, 525–8.CrossRefGoogle ScholarPubMed
Lee, D. H., Severin, K., Yokobayashi, Y., and Ghadiri, M. R. (1997). Emergence of symbiosis in peptide self-replication through a hypercyclic network. Nature, 390, 591–4.CrossRefGoogle ScholarPubMed
Leman, L., Orgel, L., and Ghadiri, M. R. (2004). Carbonyl sulfide-mediated prebiotic formation of peptides. Science, 306, 283–6.CrossRefGoogle ScholarPubMed
Leser, M. E. and Luisi, P. L. (1989). Liquid 3-phase micellar extraction of peptides. Biotech. Techniques, 3, 149–54.CrossRefGoogle Scholar
Leser, M. E. and Luisi, P. L. (1990). Application of reverse micelles for the extraction of amino acids and proteins. Chimia, 44, 270–82.Google Scholar
Levashov, A. V., Klyachko, N. L., Psbezhetski, A. V., et al. (1989). Biochim. Biophys. Acta, 988, 221–56.
Levy, M. and Ellington, A. D. (2003). Peptide-template nucleic acid ligation, J. Mol. Evol., 56, 607–15.CrossRefGoogle Scholar
Lewontin, R. C. (1993). The Doctrine of DNA – Biology as an Ideology. Penguin Books.Google Scholar
Li, T. and Nicolaou, K. C. (1994). Chemical self-replication of palindromic duplex DNA. Nature, 369, 218–21.CrossRefGoogle ScholarPubMed
Li, Y., Zhao, Y., Hatfield, S.et al. (2000). Dipeptide seryl-histidine and related oligopeptides cleave DNA, protein, and a carboxyl ester. Bioorg. Med. Chem., 12, 2675–80.CrossRefGoogle Scholar
Li, X. and Chmielewski, J. (2003). Peptide self-replication enhanced by a proline kink. J. Am. Chem. Soc., 125, 11820–1.CrossRefGoogle ScholarPubMed
Lifson, S. (1997). On the crucial stages in the origin of animate matter. J. Mol. Evol., 44, 1–8.CrossRefGoogle ScholarPubMed
Limtrakul, J., Kokpol, S., and Rode, B. M. (1985). Quantum chemical investigation on ion-dipeptide complex-formation. J. Sci. Soc. Thailand, 11, 129–33.CrossRefGoogle Scholar
Lindblom, G. and Rilfors, L. (1989). Cubic phases and isotropic structures formed by membrane lipids – possible biological relevance. Biochem. Biophys. Acta, 988, 221–56.Google Scholar
Lindsey, J. S. (1991). Self-assembly in synthetic routes to molecular devices – biological principles and chemical perspectives – a review. New J. Chem., 15, 153–80.Google Scholar
Livio, M. and Rees, J. M. (2005). Anthropic reasonings. Science, 309, 1922–3.CrossRefGoogle Scholar
Lonchin, S., Luisi, P. L., Walde, P., and Robinson, B. H. (1999). A matrix effect in mixed phospholipid/fatty acid vesicle formation. J. Phys. Chem. B, 103, 10910–16.CrossRefGoogle Scholar
Love, S. G. and Brownlee, D. E. (1993). A direct measurement of the terrestrial mass accretion rate of cosmic dust. Science, 262, 550–3.CrossRefGoogle ScholarPubMed
Lovelock, J. E. (1979). Gaia: A New Look at Life on Earth. Oxford University Press.Google Scholar
Lovelock, J. E. (1988). The Ages of Gaia. W. W. Norton & Co.Google Scholar
Luci, P. (2003). Gene cloning expression and purification of membrane proteins. ETH-Z Dissertation Nr. 15108, Zurich.Google Scholar
Luhmann, K. (1984). Soziale Systeme. Suhrkamp.Google Scholar
Luisi, P. L. (1979). Why are enzymes macromolecules?Naturwissenschaften, 66, 498–504.CrossRefGoogle ScholarPubMed
Luisi, P. L. (1985). Enzyme hosted in reverse micelles in hydrocarbon solution. Angew. Chem., 24, 439–50.CrossRefGoogle Scholar
Luisi, P. L. (1993). Defining the transition to life: self-replicating bounded structures and chemical autopoiesis. In Thinking About Biology, SFI Studies in the Sciences of Complexity, ed. Stein, W. and Varela, F. J.. Addison-Wesley-Longman.Google Scholar
Luisi, P. L. (1994). In Self-Reproduction of Supramolecular Structures, Proceedings from the Maratea Symposium, eds. Fleischacker, G., Colonna, S. and Luisi, P. L.. Kluwer.CrossRefGoogle Scholar
Luisi, P. L. (1996). Self-reproduction of micelles and vesicles: models for the mechanisms of life from the perspective of compartmented chemistry. Adv. Chem. Phys., 92, 425–38.Google Scholar
Luisi, P. L. (1997a). Die Frage nach dem Ursprung des Lebens. Schweiz. Techn. Z., 12, 10–14.Google Scholar
Luisi, P. L. (1997b). Self-reproduction of chemical structures and the question of the transition to life. In Astronomical and Biochemical Origins and the Search for Life in the Universe, eds. Cosmovici, C. B., Bowyer, S., and Werthimer, D.. Editrice Compositori, pp. 461–8.Google Scholar
Luisi, P. L. and Laane, C. (1986). Solubilization of enzymes in apolar solvents via reverse micelles. Trends Biotechnol., 4, 29–38.CrossRefGoogle Scholar
Luisi, P. L. and Laane, C. (1998). About various definitions of life. Orig. Life Evol. Biosph., 28, 613–22.CrossRefGoogle ScholarPubMed
Luisi, P. L. and Laane, C. (2001). Are micelles and vesicles chemical equilibrium systems?J. Chem. Educ., 78, 380–4.CrossRefGoogle Scholar
Luisi, P. L. and Laane, C. (2002a). Toward the engineering of minimal living cells. Anat. Rec., 268, 208–14.CrossRefGoogle Scholar
Luisi, P. L. and Laane, C. (2002b). Emergence in chemistry: chemistry as the embodiment of emergence. Found. Chem., 4, 183–200.CrossRefGoogle Scholar
Luisi, P. L. and Laane, C. (2003a). Contingency and determinism. Phil. Trans. R. Soc. Lond., A, 361, 1141–7.CrossRefGoogle Scholar
Luisi, P. L. and Laane, C. (2003b). Autopoiesis: a review and reappraisal. Naturwissenschaften, 90, 49–59.Google Scholar
Luisi, P. L. and Magid, L. (1986). Solubilization of enzymes and nucleic acids in hydrocarbon micelar solutions. Crit. Rev. Biochem., 20, 409–74.CrossRefGoogle Scholar
Luisi, P. L. and Oberholzer, T. (2001). Origin of life on Earth: molecular biology in liposomes as an approach to the minimal cell. In The Bridge between the Big Bang and Biology, ed. Giovanelli, F.. CNR Press, pp. 345–55.Google Scholar
Luisi, P. L. and Straub, B., eds. (1984). Reverse Micelles. Plenum Press.CrossRefGoogle Scholar
Luisi, P. L. and Varela, F. J. (1990). Self-replicating micelles – a chemical version of minimal autopoietic systems. Orig. Life Evol. Biosph., 19, 633–43.CrossRefGoogle Scholar
Luisi, P. L. and Walde, P., eds. (2000). Giant Vesicles, Perspectives in Supramolecular Chemistry. John Wiley & Sons Ltd.Google Scholar
Luisi, P. L., Ferri, F., and Stano, P. (2006). Approaches to semi-synthetic minimal cells: a review. Naturwissenschaften, 93, 1–13.CrossRefGoogle ScholarPubMed
Luisi, P. L., Lazcano, A., and Varela, F. (1996). In Defining Life: the Central Problem in Theoretical Biology, ed. Rizzotti, M.. University of Padova, pp. 149–65.Google Scholar
Luisi, P. L., Oberholzer, T., and Lazcano, A. (2002). The notion of a DNA minimal cell: a general discourse and some guidelines for an experimental approach. Helv. Chim. Acta, 85 (6), 1759–77.3.0.CO;2-7>CrossRefGoogle Scholar
Luisi, P. L., Pellegrini, A., and Walsoe, C. (1977b). Pepsin-catalyzed coupling between aromatic amino acid residues. Experientia, 33, 796.Google Scholar
Luisi, P. L., Giomini, M., Pileni, M. P., and Robinson, B. H. (1988). Reverse micelles as hosts for proteins and small molecules. Biochim. Biophys. Acta, 947, 209–46.CrossRefGoogle ScholarPubMed
Luisi, P. L., Scartazzini, R., Haering, G., and Schurtenberger, P. (1990). Organogels from water-in-oil microemulsions. Colloid Polymer Sci., 268, 356–74.CrossRefGoogle Scholar
Luisi, P. L., Stano, P., Rasi, S., and Mavelli, F. (2004). A possibile route to prebiotic vesicle reproduction, Artificial Life, 10, 297–308.CrossRefGoogle Scholar
Luisi, P. L., Bonner, F. J., Pellegrini, A., Wiget, P., and Wolf, R. (1979). Micellar solubilization of proteins in aprotic solvents and their spectroscopic properties. Helv. Chim. Acta, 62, 740–53.CrossRefGoogle Scholar
Luisi, P. L., Henninger, F., Joppich, M., Dossena, A., and Casnati, G. (1977a). Solubilization and spectroscopic properties of α-chymotrypsin in cyclohexane. Biochem. Biophys. Res. Commun., 74, 1384–9.CrossRefGoogle Scholar
Luther, A., Brandsch, R., and Kiedrowski, G. (1998). Surface promoted replication and exponential amplification. Nature, 396, 245–8.CrossRefGoogle ScholarPubMed
Luthi, P. and Luisi, P. L. (1984). Enzymatic-synthesis of hydrocarbon-soluble peptides with reverse micelles. J. Am. Chem. Soc., 106, 7285–6.CrossRefGoogle Scholar
Luzzati, V., Vargas, R., Mariani, P., Gulik, A., and Delacroix, H. (1993). Cubic phases of lipid-containing systems: elements of a theory and biological connotations. J. Mol. Biol., 229, 540–51.CrossRefGoogle ScholarPubMed
Ma, Q. G. and Remsen, E. F. (2002). Chemically induced supramolecular reorganization of triblock copolymer assemblies: Trapping of intermediate states via a shell-crosslinking methodology. Proc. Natl. Acad. Sci. USA, 99, 5058–63.CrossRefGoogle Scholar
Machy, P. and Leserman, L. (1987). Liposomes in Cell Biology and Pharmacology. London: John Libbey and Co. Ltd.Google Scholar
Madeira, V. M. C. (1977). Biochim. Biophys. Acta, 499, 202–211.CrossRef
Mader, S. S. (1996). Biology, 5th edn. W. C. Brown Publisher.Google Scholar
Maestro, M. and Luisi, P. L. (1990). A simplified thermodynamic model for protein uptake by reverse micelles. In Surfactants in Solution, ed. Mittal, K. L.. Plenum, vol. 9.Google Scholar
Margulis, L. (1993). Symbiosis in Cell Evolution. Freeman.Google Scholar
Margulis, L. and Sagan, D. (1995). What is Life?Weidenfeld and Nicholson.Google Scholar
Mariani, P., Luzzati, V., and Delacroix, H. (1988). Cubic phases of lipid-containing systems: structure analysis and biological implications. J. Mol. Biol., 204, 165–89.CrossRefGoogle ScholarPubMed
Marks-Tarlow, T., Robertson, R., and Combs, A. (2001). Varela and the Uroborus: the psychological significance of reentry. Cybernetics Human Knowing, 9, 31.Google Scholar
Marques, E. F., Regev, O., Khan, A., Miguel, M. D., and Lindman, B. (1998). Vesicle formation and general phase behavior in the catanionic mixture SDS-DDAB-water. The anionic-rich side. J. Phys. Chem. B, 102, 6746–58.CrossRefGoogle Scholar
Martinek, K. and Berezin, I. V. (1986). Dokl. Akdam. Nauk., SSSR, 289, 1271.
Martinek, K., Levashov, A. V., Pantin, V. I., and Berezin, I. V. (1978). Model of biological membranes or surface-layer (active center) of protein globules (enzymes) – reactivity of water solubilized by reversed micelles of aerosol OT in octane during neutral hydrolysis of picrylchloride. Doklady Akademii Nauk SSSR, 238, 626–9.Google Scholar
Martinek, K., Levashov, A. V., Klyachko, N. L., Pantin, V. I., and Berezin, I. V. (1981). The principles of enzyme stabilization. 6. Catalysis by water-soluble enzymes entrapped into reversed micelles of surfactants in organic solvents. Biochem. Biophys. Acta, 657, 277–95.Google ScholarPubMed
Martinek, K, Levashov, A. V., Klyachko, N., Khmelnttski, Y. L., and Berezin, I. V. (1986a). Micellar enzymology. Eur. J. Biochem., 155, 453–468.CrossRefGoogle Scholar
Mason, S. F. and Tranter, G. E. (1983). The parity violating energy difference between enantiomeric molecules. Chem. Phys. Lett., 94, 34.CrossRefGoogle Scholar
Mason, S. F. and Tranter, G. E. (1984). The parity violating energy difference between enantiomeric molecules. Mol. Phys., 53, 1091–111.CrossRefGoogle Scholar
Matsumura, S., Takahashi, T., Ueno, A., and Mihara, H. (2003). Complementary nucleobase interaction enhances peptide–peptide recognition and self-replicating catalysis. Chem. Eur. J., 9, 4829–37.CrossRefGoogle ScholarPubMed
Matthews, C. N. (1975). The origin of proteins, heteropolypeptides from hydrogen cyanide and water, Origin of Life, 6, 155–63.CrossRefGoogle Scholar
Maturana, H. and Varela, F. (1980). Autopoiesis and Cognition: The Realization of the Living. Reidel.CrossRefGoogle Scholar
Maturana, H. and Varela, F. (1998). The Tree of Knowledge, revised edn. Shambala.Google Scholar
Maturana, H., Lettvin, J., McCulloch, W., and Pitts, W. (1960). Life and cognition. Gen. Physiol., 43, 129–75.CrossRefGoogle Scholar
Mavelli, F. (2004). Theoretical investigations on autopoietic replication mechanisms. ETH-Z Dissertation Nr. 15218, Zurich.Google Scholar
Mavelli, F. and Luisi, P. L. (1996). Autopoietic self-reproducing vesicles: a simplified kinetic model. J. Phys. Chem., 100, 16600–7.CrossRefGoogle Scholar
Maynard-Smith, J. and Szathmáry, E. (1995). The Major Transitions in Evolution. Oxford University Press.Google Scholar
Maynard-Smith, J. and Szathmáry, E. (1999). The Origins of Life. Oxford.Google Scholar
Mayr, E. (1988). The limits of reductionism. Nature, 331, 475.CrossRefGoogle Scholar
McBride, J. M. and Carter, R. L. (1991). Spontaneous resolution by stirred crystallization. Angew. Chem. Int. Ed. Engl., 30, 293–5.CrossRefGoogle Scholar
McCollom, T. M., Ritter, G., and Simoneit, B. R. T. (1999). Lipid synthesis under hydrothermal conditions by Fischer-Tropsch-type reactions. Orig. Life Evol. Biosph., 29, 153–66.CrossRefGoogle ScholarPubMed
McLaughlin, B. P. (1992). The rise and fall of British emergentism. In Emergence or Reduction: Essays on the Prospects of Nonreductive Materialism, eds. Beckermann, A., Flohr, H. and Kim, J.. de Gruyter, pp. 49–3.CrossRefGoogle Scholar
Meier, C. A. (1992). Wolfang Pauli und C. G. Jung, Ein Briefwechsel. Springer Verlag.CrossRefGoogle Scholar
Menger, F. (1991). Groups of organic molecules that operate collectively. Angew. Chem. Int. Ed. Engl., 30, 1086–99.CrossRefGoogle Scholar
Merleau-Ponty, M. (1967). The Structure of Behaviour. Beacon.Google Scholar
Micura, R., Bolli, M., Windhab, N., and Eschenmoser, A. (1997). Angew. Chem. Int. Ed. Engl., 36, 870.CrossRef
Micura, R., Kudick, R., Pitsch, S., and Eschenmoser, A. (1999). Angew. Chem. Int. Ed. Engl., 38, 680.3.0.CO;2-C>CrossRef
Mill, J. S. (1872). System of Logic, 8th edn. Longmans, Green, Reader and Dyer.Google Scholar
Miller, C., Cuendet, P., and Gratzel, M. (1991). Adsorbed omega-hydroxy thiol monolayers on gold electrodes – evidence for electron-tunneling to redox species in solution. J. Phys. Chem., 95, 877–86.CrossRefGoogle Scholar
Miller, M. B. and Basler, B. L. (2001). Quorum sensing in bacteria. Ann. Rev. Microbiol., 55, 165–199.CrossRefGoogle ScholarPubMed
Miller, S. L. (1953). Production of amino acids under possible primitive Earth conditions. Science, 117, 2351–61.CrossRefGoogle ScholarPubMed
Miller, S. L. (1998). The endogenous synthesis of organic compounds. In The Molecular Origin of Life, ed. Brack, A.. Cambridge University Press.Google Scholar
Miller, S. L. and Bada, J. (1988). Submarine hot springs and the origin of life. Nature, 334, 609–11.CrossRefGoogle ScholarPubMed
Miller, S. L. and Bada, J. (1991). Extraterrestrial synthesis. Nature, 350, 388–89.CrossRefGoogle ScholarPubMed
Miller, S. L. and Lazcano, A. (1995). J. Mol. Evol., 41, 689–92.CrossRef
Miller, S. L. and Lazcano, A. (2002). Formation of the building blocks of life. In Life's Origin, The Beginning of Biological Evolution, ed. Schopf, J. W.. California University Press, pp. 100–9.Google Scholar
Miller, S. L. and Parris, M. (1964). Nature, 204, 1248–50.CrossRef
Mingers, J. (1992). The problems of social autopoiesis. Int. J. Gen. Syst., 21, 229–36.CrossRefGoogle Scholar
Mingers, J. (1995). Self-Producing Systems: Implications and Applications of Autopoiesis. Plenum.CrossRefGoogle Scholar
Mingers, J. (1997). A critical evaluation of Maturana's constructivist family therapy. Syst. Practice, 10 (2), 137–51.CrossRefGoogle Scholar
Miranda, M., Amicarelli, F., Poma, A., Ragnelli, A. M., and Arcadi, A. (1988). Biochim. Biophys. Acta, 966, 276–86.CrossRef
Monod, J. (1971). Chance and Necessity. A. A. Knopf.Google Scholar
Morgan, C. L. (1923). Emergent Evolution. William and Norgate.Google Scholar
Morigaki, K., Dallavalle, S., Walde, P., Colonna, S., and Luisi, P. L. (1997). Autopoietic self-reproduction of chiral fatty acid vesicles. J. Am. Chem. Soc., 119, 292–301.CrossRefGoogle Scholar
Morowitz, H. J. (1967). Biological self-replicating systems. Prog. Theor. Biol., 1, 35–58.CrossRefGoogle Scholar
Morowitz, H. J. (1992). Beginnings of Cellular Life. Yale University Press.Google Scholar
Morowitz, H. J., Deamer, D. W., and Smith, T. (1991). Biogenesis as an evolutionary process. J. Mol. Evol., 33, 207–8.CrossRefGoogle ScholarPubMed
Morowitz, H. J., Peterson, E., and Chang, S. (1995). The synthesis of glutamic acid in the absence of enzymes – implications for biogenesis. Orig. Life Evol. Biosph., 25, 395–9.CrossRefGoogle ScholarPubMed
Morowitz, H. J., Kostelnik, J. D., Yang, J., and Cody, G. D. (2000). The origin of intermediary metabolism. Proc. Natl. Acad. Sci. USA, 97, 7704–9.CrossRefGoogle ScholarPubMed
Mossa, G., Giulio, Di A., Dini, L., and Finazzi-Agrò, A. (1989). Biochim. Biophys. Acta, 986, 310–14.CrossRef
Müller, D., Pitch, S., and Kittaka, A. (1990). Chemie von α-aminonitriles, Helv. Chim. Acta, 73, 1410–68.CrossRefGoogle Scholar
Mushegian, A. (1999). The minimal genome concept. Curr. Opin. Genetics Develop., 9, 709–714.CrossRefGoogle ScholarPubMed
Mushegian, A. and Koonin, E. V. (1996). A minimal gene set for cellular life derived by comparison of complete bacterial genomes. Proc. Natl. Acad. Sci. USA, 93, 10268–73.CrossRefGoogle ScholarPubMed
Nagel, E. (1961). The Structure of Science. Harcourt.Google Scholar
Nakajima, T., Yabushita, Y., and Tabushi, I. (1975). Amino acid synthesis through biogenic CO2 fixation. Nature, 256, 60–1.CrossRefGoogle Scholar
Naoi, M., Naoi, M., Shimizu, T., Malviya, A. N., and Yagi, K. (1977). Permeability of amino acids into liposomes. Biochim. Biophys. Acta, 471, 305–10.CrossRefGoogle ScholarPubMed
Neumann, J. and Burks, A., eds. (1966). Theory of Self-Reproduction Automata. University of Illinois Press.Google Scholar
Nicolis, G. and Prigogine, I. (1977). Self-Organization in Nonequilibrium Systems. From Dissipative Structures to Order Through Fluctuations. Wiley.Google Scholar
Nissen, P., Hansen, J., Ban, N., Moore, P. B., and Steitz, T. A. (2000). The structural basis of ribosome activity in peptide bond synthesis. Science, 289, 920–30.CrossRefGoogle ScholarPubMed
Noireaux, V. and Libchaber, A. (2004). A vesicle bioreactor as a step toward an artificial cell assembly. Proc. Natl. Acad. Sci. USA, 101, 17669–74.CrossRefGoogle ScholarPubMed
Noireaux, V., Bar-Ziv, R., and Libchaber, A. (2003). Principles of cell-free genetic circuit assembly. Proc. Natl. Acad. Sci. USA, 100, 12672–7.CrossRefGoogle ScholarPubMed
Nomura, S. M., Tsumoto, K., Yoshikawa, K., Ourisson, G., and Nakatani, Y. (2002). Towards proto-cells: “primitive” lipid vesicles encapsulating giant DNA and its histone complex, Cell. Mol. Biol. Lett., 7, 245–6.Google ScholarPubMed
Nomura, S. M., Tsumoto, K., Hamada, T., et al. (2003). Gene expression within cell-sized lipid vesicles, Chem. Bio. Chem., 4, 1172–5.CrossRefGoogle ScholarPubMed
Nooner, D. W., Gilbert, J. M., Gelpi, E., and Oró, J. (1976). Closed system Fischer-Tropsch synthesis over meteoritic iron, iron-ore and nickel-iron alloy. Geochim. Cosmochim. Acta, 40, 915–24.CrossRefGoogle Scholar
Noyes, R. M. (1989). Some models of chemical oscillators. J. Chem. Educ., 66, 190–1.CrossRefGoogle Scholar
O'Connor, T. (1994). Emergent properties. Am. Phil. Q., 31, 91–104.Google Scholar
Oberholzer, T. and Luisi, P. L. (2002). The use of lipsomes for constructing cell models. J. Biol. Phys., 28, 733–44.CrossRefGoogle Scholar
Oberholzer, T., Albrizio, M., and Luisi, P. L. (1995a). Polymerase chain reaction in liposomes. Curr. Biol., 2, 677–82.Google Scholar
Oberholzer, T., Nierhaus, K. H., and Luisi, P. L. (1999). Protein expression in liposomes. Biochem. Biophys. Res. Commun., 261, 238–41.CrossRefGoogle Scholar
Oberholzer, T., Wick, R., Luisi, P. L., and Biebricher, C. K. (1995b). Enzymatic RNA replication in self-reproducing vesicles: an approach to a minimal cell. Biochem. Biophys. Res. Commun., 207, 250–7.CrossRefGoogle Scholar
Oie, T., Loew, G. H., Burt, S. K., and MacElroy, R. D. (1983). Quantum chemical studies of a model for peptide bond formation. 2. Role of amine catalyst in formation of formamide and water from ammonia and formic acid. J. Am. Chem. Soc., 105, 2221–7.CrossRefGoogle Scholar
Olsson, U. and Wennerstrom, H. (2002). On the ripening of vesicle dispersions. J. Phys. Chem. B, 106, 5135–8.CrossRefGoogle Scholar
Ono, N. and Ikegami, T. (2000). Self-maintenance and self-reproduction in an abstract cell model. J. Theor. Biol., 206, 243–53.CrossRefGoogle Scholar
Oparin, A. I. (1924). Proishkhozhddenie Zhisni. Moskowski Rabocii.Google Scholar
Oparin, A. I. (1938). Origin of Life. McMillan.Google Scholar
Oparin, A. I. (1953). The Origin of Life. Dover Publications.Google Scholar
Oparin, A. I. (1957). The Origin of Life on Earth, 3rd edn. Academic Press.Google Scholar
Oparin, A. I. (1961). Life: Its Nature, Origin and Development. Oliver and Boyd.Google Scholar
Oppenheim, P. and Putnam, H. (1958). The unity of science as a working hypothesis. In Minnesota Studies in the Philosphy of Science, eds. Feigl, H., Maxwell, G. and Scriven, M.. University of Minnesota Press, pp. 3–36.Google Scholar
Orgel, L. E. (1973). The Origins of Life. Wiley.Google Scholar
Orgel, L. E. (1992). Molecular replication. Nature, 358, 203–9.CrossRefGoogle ScholarPubMed
Orgel, L. E. (1994). The origin of life on the Earth. Sci. Amer., 271 (4), 53–61.CrossRefGoogle ScholarPubMed
Orgel, L. E. (1995). Unnatural selection in chemical systems. Acc. Chem. Res., 28, 109–18.CrossRefGoogle ScholarPubMed
Orgel, L. E. (1998). Polymerization on the rocks: theoretical introduction. Orig. Life Evol. Biosph., 28, 227–34.CrossRefGoogle ScholarPubMed
Orgel, L. E. (2000). Self-organizing biochemical cycles. Proc. Natl. Acad. Sci. USA, 97, 12503–7.CrossRefGoogle ScholarPubMed
Orgel, L. E. (2002). The origin of biological information. In Life's Origin, the Beginnings of Biological Evolution, ed. Schopf, J. W.. California University Press, pp. 1–38.Google Scholar
Orgel, L. E. (2003). Some consequences of the RNA world hypothesis. Orig. Life Evol. Biosph., 33, 211–18.CrossRefGoogle ScholarPubMed
Oró, J. (1960). Synthesis of adenine from ammonium cyanide. Biochem. Bioph. Res. Commun., 2, 407–12.CrossRefGoogle Scholar
Oró, J. (1961). Amino acid synthesis from hydrogen cyanide under possible primitive Earth conditions. Nature, 190, 442–3.CrossRefGoogle ScholarPubMed
Oró, J. (1994). In Early Life on Earth, Nobel Symposium n. 84, ed. Bengtson, S.. Columbia University Press, pp. 48–59.Google Scholar
Oró, J. (2002). Historical understanding of life's origin. In Life's Origin, the Beginnings of Biological Evolution, ed. Schopf, J. W.. California University Press, pp. 7–41.Google Scholar
Oró, J. and Kimball, A. P. (1961). Synthesis of purines under possible primitive Earth conditions. 1. Adenine from hydrogen cyanide. Arch. Biochem. Biophys., 94, 221–7.CrossRefGoogle Scholar
Oró, J. and Kimball, A. P. (1962). Synthesis of purines under possible primitive earth conditions. 2. Purine intermediates from hydrogen cyanide. Arch. Biochem. Biophys., 96, 293–313.CrossRefGoogle Scholar
Ousfouri, S., Stano, P., and Luisi, P. L. (2005). Condensed DNA in lipid microcompartments. J. Phys. Chem. B., 109, 19929–35.CrossRefGoogle Scholar
Ourisson, G. and Nakatani, Y. (1994). The terpenoid theory of the origin of cellular life: the evolution of terpenoids to cholesterol. Chem. Biol., 1, 11–23.CrossRefGoogle Scholar
Ourisson, G. and Nakatani, Y. (1999). Origin of cellular life: molecular foundations and new approaches. Tetrahedron, 55, 3183–90.CrossRefGoogle Scholar
Paechthorowitz, M. and Eirich, F. R. (1988). The polymerisation of amino acid adenilates on sodium montmorillonite with preadsorbed polypeptides. Orig. Life Evol. Biosph., 18, 359–87.CrossRefGoogle Scholar
Palazzo, G. and Luisi, P. L. (1992). Solubilization of ribosomes in reverse micelles. Biochem. Biophys. Res. Commun., 186, 1546–52.CrossRefGoogle ScholarPubMed
Paley, W. (1802) (other sources report 1803). Natural Theology, or Evidences of the Existence and Attributes of the Deity, Collected from the Appearances of Nature, 12th edn (1986). Lincoln-Rembrandt Publishing.CrossRefGoogle Scholar
, Palyi G., , Zucchi C., and , Caglioti L., eds. (2002). Fundamentals of life. Elsevier.Google Scholar
Pantazatos, D. P. and McDonald, R. C. (1999). Directly observed membrane fusion between oppositely charged phospholipid bilayers. Membrane Biol., 170, 27–38.CrossRefGoogle ScholarPubMed
Papahadjopoulus, D., Lopez, N., and Gabizon, A. (1989). In Liposomes in Therapy of Infectious Diseases and Cancer, eds. Lopez-Berenstein, G. and Fidler, I. J.. Alan Riss Inc., pp. 135–154.Google Scholar
Parsons, P. (1996). Dusting off panspermia. Nature, 383, 221–2.CrossRefGoogle ScholarPubMed
Paul, N. and Joyce, G. F. (2002). A self-replicating ligase ribozyme. Proc. Natl. Acad. Sci., 99, 12733–40.CrossRefGoogle ScholarPubMed
Paul, N. and Joyce, G. F. (2004). Minimal self-replicating systems. Curr. Opin. Chem. Biol., 8, 634–9.CrossRefGoogle ScholarPubMed
Penzien, K. and Schmidt, G. M. J. (1969). Reactions in chiral crystals – an absolute asymmetric synthesis. Angew. Chem. Int. Ed. Engl., 8, 608.CrossRefGoogle Scholar
Peretó, Y., Lopez-Garcia, P., and Moreira, D. (2004). Ancestral lipid biosynthesis and early membrane evolution. Trends Biochem. Sci., 29, 469–77.CrossRefGoogle ScholarPubMed
Pfammatter, N., Guadalupe, A. A., and Luisi, P. L. (1989). Solubilization and activity of yeast cells in water-in-oil microemulsion. Biochem. Biophys. Res. Commun., 161, 1244–51.CrossRefGoogle ScholarPubMed
Pfammatter, N., Hochkoppler, A., and Luisi, P. L. (1992). Solubilization and growth of Candida pseudotropicalis in water-in-oil microemulsions. Biotechnol. Bioeng., 40, 167–72.CrossRefGoogle ScholarPubMed
Pfüller, U. (1986). Mizellen, Vesikeln, Mikroemulsionen. Springer Verlag.CrossRefGoogle Scholar
Piaget, J. (1967). Biologie et connaissance. Gallimard.Google Scholar
Pietrini, A. V. and Luisi, P. L. (2002). Circular dichroic properties and average dimensions of DNA-containing reverse micellar aggregates. Biochim. Biophys. Acta, 1562, 57–62.CrossRefGoogle ScholarPubMed
Pietrini, A. V. and Luisi, P. L. (2004). Cell-free protein synthesis through solubilisate exchange in water/oil emulsion compartments. Chem. Bio. Chem, 5, 1055–1062.CrossRefGoogle ScholarPubMed
Pileni, M. P. (1981). Photoelectron transfer in reverse micelles – photo-reduction of cytochrome-c. Chem. Phys. Lett., 81, 603–5.CrossRefGoogle Scholar
Piries, N. M. (1953). Discovery, 14, 238.
Pizzarello, S. and Cronin, J. R. (2000). Non-racemic amino acids in the Murray and Murchison meteorites. Geochim. Cosmochim. Acta, 64, 329–38.CrossRefGoogle ScholarPubMed
Pizzarello, S. and Weber, A. L. (2004). Prebiotic amino acids as asymmetric catalysts. Science, 303, 1151.CrossRefGoogle ScholarPubMed
Pizzarello, S., Feng, X., Epstein, S., and Cronin, J. R. (1994). Isotopic N-lyses of nitrogenous compounds from Murchison meteorite: ammonia, amines, amino acids, and polar hydrocarbons. Geochim. Cosmochim. Acta, 58, 5579–87.CrossRefGoogle ScholarPubMed
Plankensteiner, K., Righi, A., and Rode, B. M. (2002). Glycine and diglycine as possible catalytic factors in the prebiotic evolution of peptides. Orig. Life Evol. Biosph., 32, 225–36.CrossRefGoogle ScholarPubMed
Plasson, R., Biron, J. P., Cottet, H., Commeyras, A., and Taillades, J. (2002). Kinetic study of the polymerization of alpha-amino acid N-carboxyanhydrides in aqueous solution using capillary electrophoresis. J. Chromatogr. A, 952, 239–48.CrossRefGoogle ScholarPubMed
Platt, J. R. (1961). Properties of large molecules that go beyond the properties of their chemical sub-groups. J. Theor. Biol., 1, 342–58.CrossRefGoogle ScholarPubMed
Poerksen, B. (2004). The Certainty of Uncertainty, Dialogues Introducing Constructivism. Imprint Academic.Google Scholar
Pohorille, A. and Deamer, D. (2002). Artificial cells: prospects for biotechnology. Trends Biotech., 20, 123–8.CrossRefGoogle ScholarPubMed
Pojman, J. A., Craven, R., and Leard, D. C. (1994). Oscillations and chemical waves in the physical chemistry lab. J. Chem. Educ., 71, 84–90.CrossRefGoogle Scholar
Ponce de Leon, S. and Lazcano, A. (2003). Panspermia – true or false?Lancet, 362, 406–7.CrossRefGoogle ScholarPubMed
Popa, R. (2004). Between Necessity and Probability: Searching for the Definition and Origin of Life. Springer Verlag.Google Scholar
Pope, M. T. and Muller, A. (1991). Polyoxometalate chemistry – an old field with new dimensions in several disciplines. Angew. Chem. Int. Ed. Engl., 30, 34–48.CrossRefGoogle Scholar
Portmann, M., Landau, E. M., and Luisi, P. L. (1991). Spectroscopic and rheological studies of enzymes in rigid lipidic matrices: the case of α-chymotrypsin in a lysolecithin/water cubic phase. J. Phys. Chem., 95, 8437–40.CrossRefGoogle Scholar
Pozzi, G., Birault, V., Werner, B. (1996). Single-chain polyprenyl phosphates form primitive membranes. Angew. Chem. Int. Ed. Engl., 35, 177–9.CrossRefGoogle Scholar
Prigogine, I. (1997). The End of Certainty-Time, Chaos and the New Laws of Nature. Free Press.Google Scholar
Prigogine, I. and Lefever, R. (1968). Symmetry breaking instabilities in dissipative systems. J. Chem. Phys., 48, 1695–700.CrossRefGoogle Scholar
Primas, H. (1985). Can chemistry be reduced in physics?Chem. Uns. Zeit, 19, 160.CrossRefGoogle Scholar
Primas, H. (1993). In Neue Horizonte 92/93: Ein Forum der Naturwissenschaften, ed. Fischer, E. P.. München: Piper.Google Scholar
Primas, H. (1998). Emergence in exact natural sciences. Acta Politechnica Scand. 91, 86–7.Google Scholar
Pryer, W. (1880). Die Hypothesen über den Ursprung des Lebens. Berlin.Google Scholar
Purrello, R. (2003). Lasting chiral memory. Nature Mater., 2, 216–17.CrossRefGoogle ScholarPubMed
Pyun, J., Zhou, X.-Z., Drockenmuller, E., and Hawker, C. J. (2003). Mater. Chem., 13, 2653.CrossRef
Quack, M. (2002). Angew. Chem., 41, 4618–30.CrossRef
Quack, M. and Stohner, J. (2003a). Combined multidimensional anharmonic and parity violating effects in CDBrClF. J. Chem. Phys., 119, 11228–40.CrossRefGoogle Scholar
Quack, M. and Stohner, J. (2003b). Molecular chirality and the fundamental symmetries of physics: influence of parity violation on rotovibrational frequencies and thermodynamic properties. Chirality, 15, 375–6.CrossRefGoogle Scholar
Raab, W. (1988). Ärtzliche Kosmetologie, 18, 213–24.
Radzicka, A. and Wolfenden, R. (1996). Rates of uncatalyzed peptide bond hydrolysis in neutral solution and the transition state affinities of proteases. J. Am. Chem. Soc., 118, 6105–9.CrossRefGoogle Scholar
Ramundo-Orlando, A., Arcovito, C., Palombo, A., Serafino, A. L., and Mossa, G. (1993). J. Liposome Res., 3, 717–24.CrossRef
Ramundo-Orlando, A., Mattia, F., Palombo, A., and D'Inzeo, G. (2000). Effect of low frequency, low amplitude magnetic fields on the permeability of cationic liposomes entrapping carbonic anhydrase, part II. Bioelectromagnetics, 21, 499–507.3.0.CO;2-9>CrossRefGoogle Scholar
Rao, M., Eichberg, J., and Oró, J. (1987). Synthesis of phosphatidylethanolamine under possible primitive earth conditions. J. Mol. Evol., 25, 1–6.CrossRefGoogle ScholarPubMed
Rasi, S., Mavelli, F., and Luisi, P. L. (2003). Cooperative micelle binding and matrix effect in oleate vesicle formation. J. Phys. Chem. B, 107, 14068–76.CrossRefGoogle Scholar
Rasi, S., Mavelli, F., and Luisi, P. L. (2004). Matrix effect in oleat-micelles-vesicles transformations. Orig. Life Evol. Bioph., 34, 215–24.CrossRefGoogle Scholar
Rathman, J. F. (1996). Micellar catalysis. Curr. Opin. Coll. Interf. Sci., 1, 514–518.CrossRefGoogle Scholar
Rebek, J. (1994). A template for life. Chem. Br., 30, 286–90.Google Scholar
Reichenbach, H. (1978). The aims and methods of physical knowledge. In Hans Reichenbach: Selected Writings 1909–53 (transl. E. H. Schneewind), eds. Reichenbach, M. and Cohen, R. S.. Reidel, pp. 81–225.Google Scholar
Reszka, R. (1998). Liposomes as drug carrier for diagnostics, cytostatics and genetic material. In Future Strategies for Drug Delivery with Particulate Systems, eds. Diederichs, J. E. and Müller, R. H.. Medpharm GmbH Scientific Publishers.Google Scholar
Ribo, J. M., Crusats, J., Sagues, F., Claret, J., and Rubires, R. (2001). Chiral sign induction during the formation of mesophases in stirred solutions. Science, 292, 2063–6.CrossRefGoogle ScholarPubMed
Rikken, G. L. and Raupach, E. (2000). Enantioselective magnetochiral photochemistry. Nature, 405, 895–6.CrossRefGoogle ScholarPubMed
Rispens, T. and Engberts, J. B. F. N. (2001). Efficient catalysis of a Diels-Alder reaction by metallo-vesicles in aqueous solution. Org. Lett., 3, 941–943.CrossRefGoogle ScholarPubMed
Riste, T. and Sherrington, D., ed. (1996). Physics of Biomaterials: Fluctuations, Selfassembly and Evolution (Nato Science Series Series E, Applied Sciences). Kluwer.CrossRefGoogle Scholar
Rizzotti, M., ed. (1996). Defining Life. University of Padua.Google Scholar
Robertson, R. N. (1983). The Lively Membrane. Cambridge University Press.Google Scholar
Rode, B. M., Son, H. L. and Suwannachot, Y. (1999). The combination of salt induced peptide formation reaction and clay catalysis: a way to higher peptides under primitive earth conditions. Orig. Life. Evol. Biosph., 29, 273–86.CrossRefGoogle ScholarPubMed
Rolle, F. (1863). Ch. Darwin's Lehre von der Entstehung der Arten, in ihrer Anwendung auf die Schöpfunggeschichte. J. C. Hermann.Google Scholar
Roseman, A., Lentz, B. R., Sears, B., Gibbes, D., and Thompson, T. E. (1978). Properties of sonicated vesicles of three synthetic phospholipids. Chem. Phys. Lipids, 21, 205–22.CrossRefGoogle Scholar
Rotello, V., Hong, J. I., and Rebek, J. (1991). Sigmoidal growth in a self-replicating system. J. Am. Chem. Soc., 113, 9422–3.CrossRefGoogle Scholar
Roux, A., Cappello, G., Carteaud, J., et al. (2002). A minimal system allowing tubulation with molecular motors pulling on giant liposomes. Proc. Natl. Acad. Sci., 99, 5394–9.CrossRefGoogle ScholarPubMed
Rushdi, A. I. and Simoneit, B. R. (2001). Lipid formation by aqueous Fischer-Tropf-type synthesis over a temperature range 100 to 400 ℃. Orig. Life Evol. Biosph., 31, 103–18.CrossRefGoogle Scholar
Sackmann, E. (1978). Dynamic molecular-organization in vesicles and membranes. Ber. Bunsen-Gesell. Phys. Chem., 82, 891–909.CrossRefGoogle Scholar
Sackmann, E. (1995). In Structure and Dynamics of Membranes, eds. Lipowsky, R. and Sackmann, E.. Elsevier Science, vol 1, pp. 213–304.Google Scholar
Sada, E., Katoh, S., Terashima, M., and Tsukiyama, K.-I. (1988). Entrapment of an ion-dependent enzyme into reverse-phase evaporation vesicles. Biotechnol. Bioeng., 32, 826–30.CrossRefGoogle ScholarPubMed
Sada, E., Katoh, S., Terashima, M., Shiraga, H., and Miura, Y. (1990). Stability and reaction characteristics of reverse-phase evaporation vesicles (revs) as enzyme containers. Biotechnol. Bioeng., 36, 665–71.CrossRefGoogle ScholarPubMed
Saetia, S., Liedl, K. R., Eder, A. H., and Rode, B. M. (1993). Evaporation cycle experiments: a simulation of salt-induced peptide synthesis under possible prebiotic conditions. Orig. Life Evol. Biosph., 3, 167–76.CrossRefGoogle Scholar
Sagan, C. (1994). The search for extraterrestrial life. Sci. Amer., 271 (4), 71–77.CrossRefGoogle ScholarPubMed
Sagan, C. (1985). Cosmos. Ballantine Publishing.Google Scholar
Sanchez, R. A., Ferris, J. P., and Orgel, L. E. (1966). Conditions for purine synthesis: did prebiotic synthesis occur at low temperature?Science, 153, 72–3.CrossRefGoogle Scholar
Sanchez, R. A., Ferris, J. P., and Orgel, L. E. (1968). Studies in prebiotic synthesis. IV, The conversion of 4-aminoimidazole-5-carbonitrile derivatives to purines. J. Mol. Biol., 38, 121–8.CrossRefGoogle ScholarPubMed
Sankararaman, S., Menon, G. I., and Kumar, P. B. S. (2004). Self-organized pattern formation in motor-microtubule mixtures. Phys. Rev., 70, 31904–18.Google ScholarPubMed
Sato, I., Kadowaki, K., Ohgo, Y., and Soai, K. (2004). Highly enantio selective asymmetric autocatalysis induced by chiral ionic crystals of sodium chlorate and sodium bromate. J. Mol. Catal. A, 216, 209–14.CrossRefGoogle Scholar
Scartazzini, R. and Luisi, P. L. (1988). Organogels from lecithins. J. Phys. Chem., 92, 829–33.CrossRefGoogle Scholar
Schaerer, A. A. (2002). Conceptual conditions for conceiving life – a solution for grasping its principle, not mere appearances. In Fundamentals of Life, eds. Palyi, G., Zucchi, C., and Caglioti, L.. Elsevier, pp. 589–624.Google Scholar
Schmidli, P. K., Schurtenberger, P., and Luisi, P. L. (1991). Liposome-mediated enzymatic synthesis of phosphatidylcholine as an approach to self-replicating liposomes. J. Am. Chem. Soc., 113, 8127–30.CrossRefGoogle Scholar
Schopf, J. W. (1992). In The Proterozoic Atmosphere, eds. Schopf, J. W. and Klein, C.. Cambridge University Press.CrossRefGoogle Scholar
Schopf, J. W. (1993). Microfossils of the early archean apex chert: new evidence of the antiquity of life. Science, 260, 640–6.CrossRefGoogle ScholarPubMed
Schopf, J. W. (1998). In The Molecular Origin of Life, ed. Brack, A.. Cambridge University Press.CrossRefGoogle Scholar
Schopf, J. W. (2002). Life's Origin. University of California Press.Google Scholar
Schröder, J. (1998). Emergence: non-deducibility or downward causation?Phil. Q., 48, 434–52.CrossRefGoogle Scholar
Schurtenberger, P., Magid, L. J., King, S. M. and Lindner, P. (1991). Cylindrical structure and flexibility of polymerlike lecithin reverse micelles. J. Phys. Chem., 95, 4173–6.CrossRefGoogle Scholar
Schurtenberger, P., Scartazzini, R., Magid, L. J., Leser, M. E., and Luisi, P. L. (1990). Structural and dynamic properties of polymer-like reverse micelles. J. Phys. Chem., 94, 3695–701.CrossRefGoogle Scholar
Schuster, P. and Swetina, J. (1988). Stationary mutant distributions and evolutionary optimization. Bull. Math. Biol., 50, 636–60.CrossRefGoogle ScholarPubMed
Schwabe, C. (2001). The Genomic Potential Hypothesis, a Chemist's View on the Origin and Evolution of Life. Landes Bioscience.Google Scholar
Schwabe, C. (2002). Genomic potential hypothesis of evolution: a concept of biogenesis in habitable spaces of the universe. Anat. Rec., 268, 171–9.CrossRefGoogle ScholarPubMed
Schwabe, C. (2004). Chemistry and biodiversity: Darwinism, evolution, and speciation, Chem. Biodiv., 1, 1588–90.CrossRefGoogle Scholar
Schwabe, C. and Warr, G. W. (1984). A polyphyletic view of evolution. The genetic potential hypothesis. Persp Biol. Med., 27, 465–85.CrossRefGoogle ScholarPubMed
Schwartz, A. W. (1998). Origin of the RNA world. In The Molecular Origins of Life, ed. Brack, A.. Cambridge University Press.CrossRefGoogle Scholar
Schwendiger, M. G. and Rode, B. M. (1992). Investigations on the mechanism of the salt-induced peptide formation. Orig. Life Evol. Biosph., 6, 349–59.CrossRefGoogle Scholar
Scott, E. (2004). Evolution vs. Creationism: An Introduction. Greenwood Press. See scott@natcenscied.org; www.ncseweb.org.Google Scholar
Seddon, J. M. (1990). Structure of the inverted hexagonal (HII) phase, and non-lamellar phase transitions of lipids, Biochim. Biophys. Acta, 1031, 1–69.CrossRefGoogle ScholarPubMed
Seddon, J. M., Hogan, J. J., Warrender, N. A., and Pebay-Peyroula, E. (1990). Prog. Coll. Polym. Sci., 81, 189–97.CrossRef
Segre, D. and Lancet, D. (2000). Composing life. EMBO Rep., 1 (3), 217–22.CrossRefGoogle ScholarPubMed
Segre, D., Ben-Eli, D., and Lancet, D. (2000). Compositional genomes: prebiotic information transfer in mutually catalytic non-covalent assemblies. Proc. Natl. Acad. Sci. USA, 97 (8), 4112–17.CrossRefGoogle Scholar
Segre, D., Ben-Eli, D., Deamer, D., and Lancet, D. (2001). The lipid world. Orig. Life Evol. Biosph., 31, 119–45.CrossRefGoogle ScholarPubMed
Severin, K., Lee, D. H., Kennan, A. J., and Ghadiri, M. R. (1997). A synthetic peptide ligase. Nature, 16 (389), 706–9.CrossRefGoogle Scholar
Shapiro, R. (1986). Origins: a Skeptic's Guide to the Creation of Life on Earth. Summit Books.Google Scholar
Shapiro, R. (1988). Prebiotic ribose synthesis: a critical analysis. Orig. Life Evol. Biosph., 18, 71–85.CrossRefGoogle ScholarPubMed
Shapiro, R. (1995). The prebiotic role of adenine: a critical analysis. Orig. Life Evol. Biosph., 25, 83–98.CrossRefGoogle ScholarPubMed
Shen, C., Lazcano, A., and Oró, J. (1990a). The enhancement activities of histidyl-histidine in some prebiotic reactions. J. Mol. Evol., 31, 445–52.CrossRefGoogle Scholar
Shen, C., Mills, T., and Oró, J. (1990b). Prebiotic synthesis of histidyl-histidine. J. Mol. Evol., 31, 175–9.CrossRefGoogle Scholar
Shenhav, B. and Lancet, D. (2004). Prospects of a computational origin-of-life endeavor. Orig. Life Evol. Biosph., 34, 181–94.CrossRefGoogle ScholarPubMed
Shermer, M. (2003). Is the universe fine-tuned for life?Sci. Amer., Jan, 23.Google Scholar
Shimkets, L. J. (1998). Structure and sizes of genomes of the archaea and bacteria. In Bacterial Genomes: Physical Structure and Analysis, eds. Bruijn, F. J., Lupskin, J. R., and Weinstock, G. M.. Kluwer, pp. 5–11.CrossRefGoogle Scholar
Shiner, E. K., Rumbaugh, K. P., and Williams, S. C. (2005). Interkingdom signaling: deciphering the language of acyl homoserine lactones. FEMS Microbiol. Rev., 29, 935–47.CrossRefGoogle Scholar
Shostak, S. (2003). Panspermia: Spreading Life Through the Universe. The SETI Institute.Google Scholar
Sievers, D., Achilles, T., Burmeister, J., et al. (1994). In Self-Production of Supramolecular Structures, eds. Fleischacker, G., Colonna, S. and Luisi, P. L.. Kluwer Publishers.CrossRefGoogle Scholar
Sievers, D. and Kiedrowski, G. (1994). Self-replication of complemenary nucleotide-based oligomers. Nature, 369, 221–4.CrossRefGoogle Scholar
Silin, V. I., Wieder, H., Woodward, J. T.et al. (2002). The role of surface free energy on the formation of hybrid bilayer membranes. J. Am. Chem. Soc., 124, 14676–83.CrossRefGoogle ScholarPubMed
Simon, H. A. (1969). The Sciences of the Artificial. MIT Press.Google Scholar
Simpson, G. G. (1973). Added comments on “The non-prevalence of humanoids”. In Communication with Extraterrestrial Intelligence, ed. Sagan, C.. MIT Press, pp. 362–4.Google Scholar
Smith, H. O., Hutchison, C. A. III, Pfannkoch, C., and Venter, J. C. (2003). Generating a synthetic genome by whole genome assembly: phiX174 bacteriophage from synthetic oligonucleotides. Proc. Natl. Acad. Sci. USA, 100, 15440–5.CrossRefGoogle ScholarPubMed
Smith, R. S. and Iglewski, B. H. (2003). P. aeruginosa quorum sensing systems and virulence. Curr. Opin. Microbiol., 6, 56–60.CrossRefGoogle ScholarPubMed
Solomon, B. and Miller, I. R. (1976). Interaction of glucose oxidase with phospholipid vesicles. Biochim. Biophys. Acta, 455, 332–42.CrossRefGoogle ScholarPubMed
Sperry, R. W. (1986). Discussions: Macro- versus Microdeterminism. Phil. Sci., 53, 265–70.CrossRefGoogle Scholar
Spirin, A. ((1986)). Ribosome Structure and Protein Synthesis. Benjamin Cummings Publishing.Google Scholar
Stano, P., Bufali, S., Pisano, C., et al. (2004). Novel campotothecin analogue (Gimatecan)-containing liposomes prepared by the ethanol injection method. J. Lipos. Res., 14, 87–109.CrossRefGoogle Scholar
Stano, P., Wehrli, E., and Luisi, P. L. (in press). Insights on the self-reproduction of oleate vesicles, J. Phys. Condensed Matter.Google Scholar
Stetter, K. O. (1998). Hyperthermophiles and their possible role as ancestors of modern life. In The Molecular Origin of Life, ed. Brack, A.. Cambridge University Press.CrossRefGoogle Scholar
Stocks, P. G. and Schwarz, A. W. (1982). Basic nitrogen-heterocyclic compounds in the Murchison meteorite. Geochim. Cosmochim. Acta, 46, 309–15.CrossRefGoogle Scholar
Strogatz, S. H. (1994). Non Linear Dynamics and Chaos, With Applications to Physics, Biology, Chemistry, and Engineering. Perseus Book Group.Google Scholar
Stryer, L. (1975). Biochemistry, Freeman and Co.Google Scholar
Suttle, D. P. and Ravel, J. M. (1974). The effects of initiation factor 3 on the formation of 30S initiation complexes with synthetic and natural messengers. Biochem. Biophys. Res. Commun., 57, 386–93.CrossRefGoogle ScholarPubMed
Suwannachot, Y. and Rode, B. M. (1999). Mutual amino acid catalysis in salt-induced peptide formation supports this mechanism's role in prebiotic peptide evolution. Origin Life Evol. Biosph., 5, 463–71.CrossRefGoogle Scholar
Swairjo, M. A., Seaton, B. A., and Roberts, M. F. (1994). Biochem. Biophys. Acta, 1191, 354–61.CrossRef
Szathmáry, E. (2002). Units of evolution and units of life. In Fundamentals of Life, eds. Palyi, G., Zucchi, L. and Caglioti, L.. Elsevier SAS, pp. 181–95.Google Scholar
Szostak, J. W., Bartel, D. P., and Luisi, P. L. (2001). Synthesizing life. Nature, 409, 387–90.CrossRefGoogle ScholarPubMed
Taillades, J., Cottet, H., Garrel, L., et al. (1999). N-Carbamoyl amino acid solid–gas nitrosation by NO/NOx: a new route to oligopeptides via α-amino acid N-carboxyanhydride. Prebiotic implications. J. Mol. Evol., 48, 638–45.CrossRefGoogle ScholarPubMed
Takahashi, Y. and Mihara, H. (2004). Construction of a chemically and conformationally self-replicating system of amyloid-like fibrils. Bioorg. Med. Chem., 12, 693–9.CrossRefGoogle ScholarPubMed
Takakura, K., Toyota, T., and Sugawara, T. (2003). A novel system of self-reproducing giant vesicles. J. Am. Chem. Soc., 125, 8134–8140. See also: Takakura, K., Toyota, T., Yamada, K., et al. (2002). Morphological change of giant vesicles triggered by dehydrocondensation reaction. ChemLett., 31, 404–5.CrossRefGoogle ScholarPubMed
Tanford, C. (1978). The hydrophobic effect and the organization of living matter. Science, 200, 1012–18.CrossRefGoogle ScholarPubMed
Teramoto, N., Imanishi, Y., and Yoshihiro, I. (2000). In vitro selection of a ligase ribozyme carrying alkylamino groups in the side chains. Bioconjugate Chem., 11, 744–8.CrossRefGoogle ScholarPubMed
Theng, B. K. G. (1974). The Chemistry of Clay-Organic Reactions. Adam Hilger.Google Scholar
Thomas, C. F. and Luisi, P. L. (2005). RNA selectively interacts with vesicles depending on their size. J. Phys. Chem. B., 109, 14544–50.CrossRefGoogle ScholarPubMed
Thomas, C. F. and Luisi, P. L. (2004). Novel properties of DDAB: matrix effect and interaction with oleate. J. Phys. Chem. B, 108, 11285–90.CrossRefGoogle Scholar
Thomas, P. J., Chiba, C. F., and McKay, C. P., eds. (1997a). Comets and the Origins and Evolution of Life. Springer Verlag.CrossRefGoogle Scholar
Thomas, R. M., Wendt, H., Zampieri, A., and Bosshard, H. R. (1995). Alpha-helical coiled coils: simple models for self-associating peptide and protein systems. Prog. Coll. Polym. Sci., 99, 24–30.CrossRefGoogle Scholar
Thomas, R. M., Zampieri, A., Jumel, K., and Harding, S. E. (1997b). A trimeric, alpha-helical, coiled coil peptide: association stoichiometry and interaction strength by analytical ultracentrifugation. Eur. Biophys. J., 25, 405–10.CrossRefGoogle Scholar
Thompson, E. and Varela, F. J. (2001a). Radical embodiment: neural dynamics and consciousness. Trends Cog. Sci., 5, 418–25.CrossRefGoogle Scholar
Tjivikua, T., Ballister, P., and Rebek, J. (1990). A self-replicating system. J. Am. Chem. Soc., 112, 1249–50.CrossRefGoogle Scholar
Tranter, G. E. (1985a). The parity-violating energy difference between enantiomeric reactions. Chem. Phys. Lett., 115, 286.CrossRefGoogle Scholar
Tranter, G. E. (1985b). The effects of parity violation on molecular structure. Chem. Phys. Lett., 121, 339.CrossRefGoogle Scholar
Tranter, G. E., MacDermott, A. J., Overill, R. E., and Speers, P. (1992). Computational studies of the electroweak origin of biomolecular handedness in natural sugars. Proc. Royal Soc., London A, 436, 603–15.CrossRefGoogle Scholar
Trinks, H., Schroeder, W., and Biebricher, C. K. (2003). Eis und die Entstehung des Lebens (Ice and the Origin of Life). Shaker Verlag.Google Scholar
Tsukahara, H., Imai, E. I., Honda, H., Hatori, K., and Matsuno, K. (2002). Prebiotic oligomerization on or inside lipid vesicles in hydrothermal environments. Orig. Life Evol. Biosph., 32, 13–21.CrossRefGoogle ScholarPubMed
Tsumoto, K., Nomura, S. M., Nakatani, Y., and Yoshikawa, K. (2002). Giant liposome as a biochemical reactor: transcription of DNA and transportation by laser tweezers. Langmuir, 17, 7225–8.CrossRefGoogle Scholar
Turing, A. (1952). The chemical basis of Morphogenesis. Phil. Trans. Royal. Soc. London B, 237, 37.CrossRefGoogle Scholar
Ulbricht, W. and Hoffmann, H. (1993). Physikalische Chemie der Tenside. In Die Tenside, ed. Kosswig, K., and Stache, H.. Carl Hanser Verlag, pp. 1–114.Google Scholar
Ulman, A. (1996). Formation and structure of self-assembled monolayers. Chem. Rev., 96, 1533–54.CrossRefGoogle ScholarPubMed
Uster, P. S. and Deamer, D. W. (1981). Fusion competence of phosphatidylserine-containing liposomes quantitatively measured by a fluorescence resonance energy transfer assay. Arch. Biochem. Biophys., 209 (2), 385–95.CrossRefGoogle ScholarPubMed
Vaida, M., Popovitz-Biro, R., Leiserowitz, L., and Lahav, M. (1991). Probing reaction pathways via asymmetric transformations in chiral and centrosymmetric crystals. In Photochemistry in Organized and Condensed Media, ed. Ramamurthy, V.. VCH, pp. 248–302.Google Scholar
Valenzuela, C. Y. (2002). Does biotic life exist. In Fundamentals of Life, eds. Palyi, G., Zucchi, C., and Cagiliati, L.. Elsevier, pp. 331–4.Google Scholar
Varela, F. J. (1979). Principles of Biological Autonomy. North Holland/Elsevier.Google Scholar
Varela, F. J. (1989a). Reflections on the circulation of concepts between a biology of cognition and systemic family therapy. Family Process, 28, 15–24.CrossRefGoogle Scholar
Varela, F. J. (1989b). Autonomie et Connaissance. Seuil, p. 167.Google Scholar
Varela, F. J. (1999). Ethical Know-How: Action, Wisdom, and Cognition. Stanford University Press.Google Scholar
Varela, F. J. (2000). El Fenomeno de la Vita. Dolmen Ensayo.Google Scholar
Varela, F. J., Maturana, H. R., and Uribe, R. B. (1974). Autopoiesis: the organization of living system, its characterization and a model. Biosystems, 5, 187–96.CrossRefGoogle Scholar
Varela, F. J., Thompson, E., and Rosch, E. (1991). The Embodied Mind.MIT Press.Google Scholar
Wächtershäuser, G. (1988). Before enzymes and templates: theory of surface metabolism. Microbiol. Rev. 52, 452–84.Google ScholarPubMed
Wächtershäuser, G. (1990a). Evolution of the first metabolic cycles. Proc. Natl. Acad. Sci. USA, 87, 200–4.CrossRefGoogle Scholar
Wächtershäuser, G. (1990b). The case for the chemoautotrophic origin of life in the iron–sulfur world. Origin Life Evol. Biosph., 20, 173–6.CrossRefGoogle Scholar
Wächtershäuser, G. (1992). Groundworks for an evolutionary biochemistry: the iron–sulfur world. Prog. Biophys. Mol. Biol., 58, 85–201.CrossRefGoogle Scholar
Wächtershäuser, G. (1997). The origin of life and its methodological challenge. J. Theor. Biol., 187, 483–94.CrossRefGoogle ScholarPubMed
Wächtershäuser, G. (2000). Life as we don't know it. Science, 289, 1307–8.CrossRefGoogle Scholar
Waks, M. (1986) Proteins and peptides in water-restricted environments. Proteins, 1, 4–15.CrossRefGoogle Scholar
Walde, P. (2000). Enzymatic reactions in giant vesicles. In: Giant Vesicles, Perspectives in Supramolecular Chemistry, eds. Luisi, P. L. and Walde, P.. John Wiley & Sons Ltd., pp. 297–311.Google Scholar
Walde, P. and Ichikawa, S. (2001). Enzymes inside lipid vesicles: preparation, reactivity and applications. Biomol. Eng., 18, 143–77.CrossRefGoogle ScholarPubMed
, Walde P. and , Luisi P. L. (1989). A continuous assay for lipases in reverse micelles based on fourier transform infrared spectroscopy. Biochemistry, 28, 3353–7.CrossRefGoogle Scholar
Walde, P. and Mazzetta, B. (1998) Bilayer permeability-based substrate selectivity of an enzyme in liposomes. Biotechnol. Bioeng., 57, 216–19.3.0.CO;2-E>CrossRefGoogle Scholar
Walde, P., Goto, A., Monnard, P.-A., Wessicken, M., and Luisi, P. L. (1994a). Oparin's reactions revisited: enzymatic synthesis of poly(adenylic acid) in micelles and self-reproducing vesicles. J. Am. Chem. Soc., 116, 7541–7.CrossRefGoogle Scholar
Walde, P., Wick, R., Fresta, M., Mangone, A., and Luisi, P. L. (1994b). Autopoietic self-reproduction of fatty acid vesicles. J. Am. Chem. Soc., 116, 11649–54.CrossRefGoogle Scholar
Weaver, W. (1948). Science and complexity. Amer. Sci., 36, 536–44.Google ScholarPubMed
Weber, A. (2002). The ‘surplus of meaning’. Biosemiotic aspects in Francisco J. Varela's philosophy of cognition. Cybernetics Human Knowing, 9, 11–29.Google Scholar
Weiner, A. M. and Maizels, N. (1987). tRNA-Like structures tag the 3′ ends of genomic RNA molecules for replication: implications for the origin of protein synthesis. Proc. Natl. Acad. Sci. USA, 84, 7383–7.CrossRefGoogle ScholarPubMed
Weissbuch, I., Popovitz-Biro, R., Leizerowitz, L., and Lahav, M. (1994). In The Lock-and the Key Principle The State of the Art-100 Years On, 1, ed. Behr, J.-P.. John Wiley & Sons Ltd., p. 173 and references cited therein.Google Scholar
Weissbuch, I., Frolow, F., Addadi, L., Lahav, M., and Leiserowitz, L. (1990). Oriented crystallization as a tool for detecting ordered aggregates of water-soluble hydrophobic alfa-amino acids at the air-solution interface. J. Am. Chem. Soc., 112, 7718–24.CrossRefGoogle Scholar
Weissbuch, I., Addadi, L., Berkovitch-Yellin, Z., et al. (1984). Spontaneous generation and amplification of optical activity in amino acids by enantioselective occlusion into centrosymmetric crystals of glycine. Nature, 310, 161–4.CrossRefGoogle Scholar
Weissbuch, I., Zepik, H., Bolbach, G., et al. (2003). Homochiral oligopeptides by chiral amplification within two-dimensional crystalline self-assemblies at the air–water interface; relevance to biomolecular handedness. Chemistry, 9 (8), 1782–94.CrossRefGoogle ScholarPubMed
Wendt, H., Durr, E., Thomas, R. M., Przybylski, M., and Bosshard, H. R. (1995). Characterization of leucine zipper complexes by electrospray ionization mass spectrometry. Protein Sci., 4, 1563–70.CrossRefGoogle ScholarPubMed
Wenneström, H. and Lindmann, B. (1979). Phys. Rev., 52, 1–86.
Westhof, E. and Hardy, N., eds. (2004). Folding and Self-Assembly of Biological Macromolecules. World Scientific Publishing Company Inc.CrossRefGoogle Scholar
Whitesides, G. M. and Boncheva, M. (2002). Beyond molecules: self-assembly of mesoscopic and macroscopic components. Proc. Natl. Acad. Sci. USA, 99, 4769–74.CrossRefGoogle ScholarPubMed
Whitesides, G. M., Mathias, J. P., and Seto, C. T. (1991). Molecular self-assembly and nanochemistry – a chemical strategy for the synthesis of nanostructures. Science, 254, 1312–19.CrossRefGoogle ScholarPubMed
Wick, R., Walde, P., and Luisi, P. L. (1995). Autocatalytic self-reproduction of giant vesicles. J. Am. Chem. Soc., 117, 1435–6.CrossRefGoogle Scholar
Willimann, H. and Luisi, P. L. (1991). Lecithin organogels as matrix for the transdermal transport of drugs. Biochem. Biophys. Res. Commun., 177, 897–900.CrossRefGoogle ScholarPubMed
Wilschut, J., Duzgunes, N., Fraley, R., and Papahadjopoulos, D. (1980). Studies on the mechanism of membrane-fusion – kinetics of calcium-ion induced fusion of phosphatidylserine vesicles followed by a new assay for mixing of aqueous vesicle contents. Biochemistry, 19, 6011–21.CrossRefGoogle ScholarPubMed
Wilson, T. L. (2001). The search for extraterrestrial intelligence. Nature, 409, 1110–14.CrossRefGoogle ScholarPubMed
Wimsatt, W. C. (1972). Complexity and organization. In Boston Studies in the Philosophy of Science, Proceedings of the Philosphy of Science Association, eds. Schaffner, K. F. and Cohen, R. S.. Reidel, pp. 67–86.Google Scholar
Wimsatt, W. C. (1976a). Reductionism, levels of organization, and the mind-body problem. In Consciousness and the Brain, eds. Globus, G., Maxwell, G., and Savodinik, I.. Plenum Press, pp. 205–66.CrossRefGoogle Scholar
Wimsatt, W. C. (1976b). Reductive explanation, a functional account. In Proceedings of the Meetings of the Philosophy of Science Association 1974, eds. Hooker, C. A., Pearse, G., Michealos, A. C., and Evra, J. W.Reidel, pp. 671–710.Google Scholar
Winfree, A. T. (1984). The prehistory of the Belousov-Zhabotinsky oscillator. J. Chem. Educ., 61, 661–3.CrossRefGoogle Scholar
Woese, C. R. (1979) A proposal concerning the origin of life on the planet Earth. J. Mol. Evol. 13, 95–101.CrossRefGoogle ScholarPubMed
Woese, C. R. (1983). The primary lines of descent and the universal ancestor. In Evolution from Molecules to Man, ed. Bendall, D. S.. Cambridge Universiy Press, pp. 209–33.Google Scholar
Wong, J. T.-F., Xue, H. (2002). Self-perfecting evolution of heteropolymer building blocks and sequences as the basis of life. In Fundamentals of Life, eds. Pályi, G., Zucchi, L., and Caglioti, L.. Elsevier SAS, pp. 473–94.Google Scholar
Wood, W. B. (1973). Genetic control of bacteriophage T4 morphogenesis. In Genetic Mechanisms of Development, ed. Ruddle, F. J.. Academic Press, pp. 29–46.Google Scholar
Woodle, M. C. and Lasic, D. D. (1992). Biochim. Biophys. Acta, 1113, 171–99.CrossRef
Yao, S., Ghosh, I., and Chmielewski, J. (1998). Selective amplification by auto- and cross-catalysis in a replicating peptide system. Nature, 396, 447–50.CrossRefGoogle Scholar
Yao, S., Ghosh, I., Zutshi, R., and Chmielewski, J. (1997). A pH-modulated self-replicating peptide. J. Am. Chem. Soc., 119, 10559–60.CrossRefGoogle Scholar
Yaroslavov, A. A., Udalyk, O. Y., Kabanov, V. A., and Menger, F. M. (1997). Manipulation of electric charge on vesicles by means of ionic surfactants: effects of charge on vesicle mobility, integrity, and lipid dynamics. Chem. Eur. J., 3, 690–5.CrossRefGoogle Scholar
Yoshimoto, M., Walde, P., Umakoshi, H., and Kuboi, R. (1999). Conformationally changed cytochrome c-mediated fusion of enzyme- and substrate-containing liposomes. Biotechnol. Prog., 15, 689–96.CrossRefGoogle ScholarPubMed
Yu, W., Sato, K., Wakabayashi, M., et al. (2001). Synthesis of functional protein in liposome. J. Biosc. Bioeng., 92, 590–3.CrossRefGoogle Scholar
Yuen, G. U. and Knenvolden, K. A. (1973). Monocarboxylic acids in Murray and Murchison carbonaceous meteorites. Nature, 246, 301–2.CrossRefGoogle Scholar
Yuen, G. U., , Lawless J. G., and Edelson, E. H. (1981). Quantification of monocarboxylic acids from a spark discharge synthesis. J. Mol. Evol., 17, 43–7.CrossRefGoogle Scholar
Zamarev, K. I., Romannikov, V. N., Salganik, R. I., Wlassoff, W. A., and Khramtsov, V. V. (1997). Modelling of the prebiotic synthesis of oligopeptides: silicate catalysts help to overcome the critical stage. Orig. Life Evol. Biosph., 27, 325–37.CrossRefGoogle Scholar
Zampieri, G. G., Jäckle, H., and Luisi, P. L. (1986). Determination of the structural parameters of reverse micelles after uptake of proteins. J. Phys. Chem., 90, 1849.CrossRefGoogle Scholar
Zeleny, M. (1977). Self-organization of living systems formal model of autopoiesis. Int. J. Gen. Syst., 4, 13–28.CrossRefGoogle Scholar
Zelinski, W. S. and Orgel, L. E. (1987). Autocatalytic synthesis of a tetranucleotide analogue. Nature, 327, 346–7.CrossRefGoogle Scholar
Zeng, F. W. and Zimmermann, S. C. (1997). Dendrimers in supramolecular chemistry: from molecular recognition to self-assembly. Chem. Rev., 97, 1681–712.CrossRefGoogle ScholarPubMed
Zeng, X., Ungar, G., Liu, Y., et al. (2004). Supramolecular dendritic liquid quasicrystals. Nature, 428, 157–60.CrossRefGoogle ScholarPubMed
Zepik, H. H., Bloechliger, E., and Luisi, P. L. (2001). A chemical model of homeostasis. Angew. Chem. Int. Ed. Engl., 40, 199–202.3.0.CO;2-H>CrossRefGoogle ScholarPubMed
Zhang, B. and Cech, T. R. (1998). Peptidyl-transferase ribozymes: trans reactions, structural characterization and ribosomal RNA-like features. Chem. Biol., 5, 539–53.CrossRefGoogle ScholarPubMed
Zhao, M. and Bada, J. L. (1989). Extraterrestrial amino acids in cretaceous/tertiary boundary sediments at Stevns Klint, Denmark. Nature, 339, 463–5.CrossRefGoogle ScholarPubMed
Zhu, J., Zhang, L., and Reszka, R. (1996a). Liposome-mediated delivery of genes and oligonucleotides for the treatment of brain tumors. In Targeting of Drugs: Strategies for Oligonucleide and Gene Delivery in Therapy, eds. Gregoriadis, G. and McCormack, B.. Plenum Press, pp. 169–87.CrossRefGoogle Scholar
Zhu, J., Zhang, L., Hanisch, U. K., Felgner, P. L., and Reszka, R. (1996b). In vivo gene therapy of experimental brain tumors by continuous administration of DNA-liposome complexes. Gene Therapy, 3, 472–6.Google Scholar
Ziegler, M., Davis, A. V., Johnson, D. W., and Raymond, K. N. (2003). Supramolecular chirality: a reporter of structural memory. Angew. Chem. Int. Ed. Engl., 42, 665–8.CrossRefGoogle ScholarPubMed
Zimmer, C. (2003). Tinker, tailor: can Venter stitch together a genome from scratch?Science, 299, 1006–8.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Pier Luigi Luisi, ETH Zentrum, Switzerland
  • Book: The Emergence of Life
  • Online publication: 17 December 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511817540.014
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Pier Luigi Luisi, ETH Zentrum, Switzerland
  • Book: The Emergence of Life
  • Online publication: 17 December 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511817540.014
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Pier Luigi Luisi, ETH Zentrum, Switzerland
  • Book: The Emergence of Life
  • Online publication: 17 December 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511817540.014
Available formats
×