Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-27T02:01:34.454Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 September 2016

Pier Luigi Luisi
Affiliation:
Università degli Studi Roma Tre
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. 419 - 456
Publisher: Cambridge University Press
Print publication year: 2016

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 Palyi, G., Zucchi, C., and Caglioti, L., eds., Fundamentals of Life. Elsevier, pp. 57–72.
Achilles, T. and von Kiedrowski, G. (1993). A self-replicating system from three starting materials. Angew. Chem., 32, 1198–11201.Google Scholar
Adamala, K. and Szostak, J. W. (2013). Nonenzymatic template-directed RNA synthesis inside model protocells. Science, 342, 1098.Google Scholar
Addiscott, T. (2011). Emergence or self-organization? Look to the soil population. Commun. Integr Biol., 4(4): 469–470.Google Scholar
Aguilar, A. (2009). What is Death? A scientific, philosophical and theological exploration of life's end. Ateneo Pontificio Apostolorum.
Ajikumar, P. K., Xiao, W. H., Tyo, K. E., Wang, Y., Simeon, F., Leonard, E., Mucha, O., Phon, T. H., Pfeifer, B., Stephanopoulos, G.(2010). Isoprenoid pathway optimization for taxol precursor overproduction in Escherichia coli. Science, 330(6000): 70–74.Google Scholar
Akanuma, S., Kigawa, T., and Yokoyama, S. (2002). Combinatorial mutagenesis to restricted amino acid usage in an enzyme to a reduced set. Proc. Natl. Acad. Sci. USA, 99, 13549–13553.Google Scholar
Alberts, B., Bray, D., Lewis, J., et al. (1989). Molecular Biology of the Cell, edn. New York: Garland Publications.
Alberts, B., Johnson, A., Lewis, J., et al. (2002). Molecular Biology of the Cell, edn. New York: Garland Publications.
Alberts, B., Johnson, A., Lewis, J., et al. (2007). Molecular Biology of the Cell, edn. New York: Garland Publications.
Alexander, S. (1920). Space, Time, and Deity. London: Mamillan.
Allison, A. C. and Gregoriadis, G. (1974). Liposomes as immunological adjuvants. Nature, 252, 252–258.Google Scholar
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, 1878–1887.Google Scholar
Anastasi, C., Buchet, F. F., Crowe, M. A., Parkes, A. L., Powner, M. W., Smith, J. M., and Sutherland, J. D. (2007). RNA: prebiotic product, or biotic invention? Chem. Biodivers., 4(4): 721–739.Google Scholar
Anderson, G. and Luisi, P. L. (1979). Papain-induced oligomerization of alpha amino acid esters. Helv. Chim. Acta, 62, 488–494.Google Scholar
Anella, F. (2011). Structural and functional exploration of the RNA sequence space. Implications for the origin of life and biotechnology. Ph.D. thesis, University Roma Tre.
Anella, F., Chiarabelli, C., De Lucrezia, D., and Luisi, P. L. (2011). Stability studies on random folded RNAs (“never born RNAs”), implications for the RNA world. Chemistry & Biodiversity, 8, 1422–1432.Google Scholar
Anfinsen, C. B. and Haber, E. (1961). Studies on the reduction and re-formation of protein disulfide bonds. J. Biol. Chem, 236, 1361–1363.Google Scholar
Anfinsen, C. B., Haber, E., Sela, M., and White, F. H. Jr. (1961). The kinetics of formation of native ribonuclease during oxidation of the reduced polypeptide chain. PNAS, 47, 1309–1314.Google Scholar
Angelova, M. I. and Dimitrov, D. S. (1988). A mechanism of liposome electro-formation. Progr. Colloid Polymer. Sci., 76, 59–67.Google Scholar
Annaluru, N., Muller, H., Mitchell, L. A., et al. (2014). Total synthesis of a functional designer eukaryotic chromosome. Science, 344(6179): 55–58.Google Scholar
Annesini, M. C., Di Giulio, A., Di Marzio, L., Finazzi-Agrò, A., and Mossa, G. (1992). J. Liposome Res., 2, 455–467.
Annesini, M. C., Di Giorgio, L., Di Marzio, L., et al. (1993). J. Liposome Res., 3, 639–48.
Annesini, M. C., Di Marzio, 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 Scholar
Apte, P. (2002). Vedantic view of life. In Palyi, G., Zucchi, C., and Caglioti, L., eds., Fundamentals of Life. Elsevier, pp. 497–502.
Archibald, J. M. (2009). The puzzle of plastid evolution. Current Biology, 19(2): R81–R88.Google Scholar
Arinin, E. I. (2002). Essence of organic life in Russian orthodox and modern philosophical tradition: beyond functionalism and elementarism. In Palyi, G., Zucchi, C., and Caglioti, L., eds., Fundamentals of Life. Elsevier, pp. 503–516.
Ashkenasy, G., Jagasia, R., Yadav, M., and Ghadiri, M. R. (2004). Design of a directed molecular network. Proc. Natl. Acad. Sci., 101, 10872–10877.Google Scholar
Atmanspacher, H. and Bishop, R. (2002). Between Chance and Choice, Interdisciplinary Perspectives on Determinism. Imprint Academic.
Atsumi, S. and Liao, J. C. (2008). Metabolic engineering for advanced biofuels production from Escherichia coli. Curr Opin Biotechnol., 19(5): 414–419.Google Scholar
Avetisov, V. V. and Goldanskii, V. I. (1991). Homochirality and stereospecific activity: evolutionary aspects. Biosystems, 25(3): 141–149.Google Scholar
Ayala, F. J. (1983). Beyond Darwinism? The challenge of macroevolution to the synthetic theory of evolution. In Asquith, P. D. and Nickles, T., eds., PSA 1982: Proceedings of the 1982 Biennial Meeting of the Philosophy of Science Association Symposia, Vol. 2, pp. 275–292.Google Scholar
Baas, N. A. (1994). Emergence, hierarchies, and hyperstructures. In Langton, C. G., ed., Artificial Life III, Santa Fe Studies in the Science of Complexity, Vol. XVII. Addison-Wesley, pp. 515–537.
Bachmann, P. A. (1991). Self-replicating micelles: aqueous micelles and enzymatically driven reactions in reverse micelles. J. Am. Chem. Soc., 113, 8204–8209.Google 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, 8200–8201.Google Scholar
Bachmann, P. A., Luisi, P. L., and Lang, J. (1992). Autocatalytic self-replication of micelles as models for prebiotic structures. Nature, 357, 57–59.Google Scholar
Bada, J. L. (1997). Meteoritics – extraterrestrial handedness? Science, 275, 942–943.Google Scholar
Bada, J. L. and Lazcano, A. (2002). Some like it hot, but not the first biomolecules. Science, 296, 1982–1983.Google Scholar
Bada, J. L. and Lazcano, A. (2003). Prebiotic soup – revisiting the Miller experiment. Science, 300, 745–746.Google Scholar
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, 453–461.Google Scholar
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, 560–568.Google Scholar
Bain, A. (1870). Logic, Books II and III. Longmans, Green & Co.
Bak, P., Tang, C., and Wiesenfeld, K. (1987). Self-organized criticality: an explanation of 1/f noise. Physical Review Letters, 59(4): 381–384.Google Scholar
Bak, P., Tang, C., and Wiesenfeld, K. (1988). Self-organized criticality. Physical Rev. A., 38, 364–374.Google Scholar
Ballard, D. G. H. and Bamford, C. H. (1956). Studies in polymerization. X. “The chain-effect.” R. Soc. Lond. A, 236 (1206): 384–396.Google Scholar
Barabási, A.-L. and Réka, A. (1999). Emergence of scaling in random networks. Science, 286 (5439): 509–512.Google Scholar
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, 4239–4244.Google Scholar
Barrow, J. D. (2001). Cosmology, life and the anthropic principle. Ann. NY Acad. Sci., 950, 139–153.Google Scholar
Barrow, J. D. and Tipler, F. J. (1986). The Anthropic Cosmological Principle. Oxford University Press.
Barrow, J. D. and Tipler, F. J. (1988). Action principles in nature. Nature, 331, 31–34.Google Scholar
Bartel, D. P. and Szostak, J. W. (1993). Isolation of new ribozymes from a large pool of random sequences. Science, 261, 1411–1418.Google Scholar
Bashan, A., Belousoff, M. J., Davidovich, C., and Yonath, A. (2010). Linking the RNA world to modern life: The proto-ribosome conception. Orig. Life Evol. Biosph., 40, 425–429.Google Scholar
Bedau, M. A. (1997). Weak emergence. In Tomberlin, J., ed., Philosophical Perspectives: Mind, Causation and World, Vol. 11. Malden, MA: Blackwell, pp. 375–399.
Beer, S. (1980). Preface. In Maturana, H. and Varela, F. J., Autopoiesis and Cognition (see infra).
Bell, E. A., Boehnke, P., Harrison, T. M., and Mao, W. L. (2015). Potentially biogenic carbon preserved in a 4.1 billion-year-old zircon. Proc. Natl. Acad. Sci. USA, 112(47): 14518–14521.Google Scholar
Belousoff, M. J., Davidovich, C., Bashan, A., and Yonath, A. (2010). On the development towards the modern world: a plausible role of uncoded peptides in the RNA world. In Ruiz-Mirazo, K. and Luisi, P. L., eds., Origins of life and evolution of biospheres, 40 (Special Issue 4–5): pp. 415–419.Google Scholar
Ben Jacob, E., Becker, I., Shapira, Y., and Levine, H. (2004). Bacterial linguistic communication and social intelligence. Trends Microbiol., 12, 366–72.Google Scholar
Benner, S. A. and Sismour, A. M. (2005). Synthetic biology. Nature Rev. Gen., 6, 524–45.Google Scholar
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, 1065–1071.Google 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, 1056–1064.Google Scholar
Bernal, J. D. (1951). The Physical Basis of Life. Routledge & Paul.
Bernal, J. D. (1965). Molecular structure, biochemical function, and evolution. In Waterman, T. H. and Morowitz, H. J., eds., Theoretical and Mathematical Biology. Blaisdell.
Bernal, J. D. (1967). The Origin of Life. World Publishing Company.
Bernal, J. D. (1971). Der Ursprung des Lebens. Editions Rencontre.
Bernard, C. (1865). Introduction to the Study of Experimental Medicine. Translated by Greene, H. C.,1927. Henry Schuman.
Bernhardt, H. S. (2012). The RNA world hypothesis: the worst theory of the early evolution of life (except for all the others). Biol Direct, 7, 23.Google 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, 303–338.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.Google 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, 273–283.Google 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.Google Scholar
Birdi, K. S. (1999). Self-Assembly Monolayer Structures of Lipids and Macromolecules at Interfaces. Plenum Press.
Biron, J.-Ph. and Pascal, R. (2004). Amino acid N-Carboxyanhydrides: activated peptide monomers behaving as phosphate-activating agents in aqueous solution. J. Am. Chem. Soc., Aug., 126(30): 9198–9199.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.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.Google Scholar
Bitbol, M. and Luisi, P. L. (2011). Science and the self-referentiality of consciousness. In Conciousness and the Universe. Cambridge, MA: Cosmology Science.
Blacka, R. A., Blosser, M. C., Stottrup, B. L., et al. (2013). Nucleobases bind to and stabilize aggregates of a prebiotic amphiphile, providing a viable mechanism for the emergence of protocells. PNAS, 110 (3): 13272–13276.Google Scholar
Blain, J. C. and Szostak, J. W. (2014). Progress toward synthetic cells. Annu. Rev. Biochem., 83, 11.1–11.26.Google Scholar
Blain, J. C., Ricardo, A., and Szostak, J. W. (2014). Synthesis and nonenzymatic template-directed polymerization of 2’-amino-2’-deoxythreose nucleotides. J. Am. Chem Soc., 136, 2033–2039.Google 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.Google 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, 5787–5796.Google Scholar
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, 842–848.Google Scholar
Blöchliger, E., Blocher, M., Walde, P., and Luisi, P. L. (1998). Matrix effect in the size distribution of fatty acid vesicles. J. Phys. Chem., 102, 10383–10390.Google Scholar
Böhler, C., Bannwarth, W., and Luisi, P. L. (1993). Self-replication of oligonucleotides in reverse micelles. Helv. Chim. Acta, 76, 2313–2320.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, 490–496.Google Scholar
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 Palyi, G., Zucchi, C., and Caglioti, L., eds., Fundamentals of Life. Elsevier, pp. 211–218.
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, 309–320.Google Scholar
Bolli, M., Micura, R., Pitsch, S., and Eschenmoser, A. (1997b). Pyranosyl-RNA: further observations on replication. Helv. Chim. Acta, 80, 1901–1951.Google 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, 6661–6663.Google Scholar
Bonaccio, S., Cescato, C., Walde, P., and Luisi, P. L. (1994b). Self-production of supramolecular structures. In Fleischaker, G. R. et al., eds., Liposomes from Lipidonucleotides and from Lipidopeptides. Kluwer Academic, pp. 225–259.
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, 4976–4978.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, 1952–1956.Google Scholar
Böttcher, B., Lucken, U., and Graber, P. (1995). The structure of the H+-ATPase from chloroplasts by electron cryomicroscopy. Biochem. Soc. Trans., 23, 780–785.Google Scholar
Bourgine, P. and Stewart, J. (2004). Autopoiesis and cognition. Artificial Life, 10(3): 327–345.Google Scholar
Bozic, B. and Svetina, S. (2004). A relationship between membrane properties forms the basis of a selectivity mechanism for vesicle self-reproduction. Eur. Bioph. J., 33, 565–571.Google Scholar
Brack, A. (ed.) (1998). The Molecular Origin of Life. Cambridge University Press.
Brasier, M. D., Green, O. R., Jephcoat, A. P., et al. (2002). Questioning the evidence for Earth's oldest fossils. Nature, 416, 76–77.Google Scholar
Briggs, T. and Rauscher, W. (1973). An oscillating iodine clock. J. Chem. Educ., 50, 496.Google 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.
Brunner, J., Graham, D. E., Hauser, H., and Semenza, G. (1980). Ion and sugar permeabilities of lecithin bilayers: Comparison of curved and planar bilayers. J. Membr. Biol., 57, 133–141.Google Scholar
Buchet, F. F. and Sutherland, J. D. (2006). Synthesis of pyrimidic nucleotides under potentially prebiotic conditions. Origins of Life and Evolution of the Biosphere, 36, 259.Google Scholar
Bucknall, D. G. and Anderson, H. L. (2003). Polymers get organized. Science, 302, 1904–1905.Google 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, 3910–3917.Google Scholar
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, 10552–10559.Google 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, 1033–1039.Google 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, 431–441.Google Scholar
Burmeister, J. (1998). Self-replication and autocatalysis. In Brack, A., ed., The Molecular Origin of Life. Cambridge University Press, pp. 295–310.
Butler, P. J. (1999). Self-assembly of tobacco mosaic virus: the role of an intermediate aggregate in generating both specificity and speed. Phil. Trans. R. Soc. Lond., 354(1383): 537–550.Google Scholar
Caffrey, M. (2015). A comprehensive review of the lipid cubic phase or in meso method for crystallizing membrane and soluble proteins and complexes. Acta Crystallogr. F. Struct. Biol. Commun., 71, 3–18.Google Scholar
Cairns-Smith, A. G. (1977). Takeover mechanisms and early biochemical evolution. Biosystems, 9, 105–109.Google Scholar
Cairns-Smith, A. G. (1978). Precambrian solution photochemistry, inverse segregation, and banded iron formations. Nature, 276, 808–809.Google Scholar
Cairns-Smith, A. G. (1982). Genetic Takeover and the Mineral Origins of Life. Cambridge University Press.
Cairns-Smith, A. G. (1985). Seven Clues to the Origin of Life. Cambridge University Press.
Cairns-Smith, A. G. (1990). Seven Clues to the Origin of Life, edn. Cambridge University Press.
Cairns-Smith, A. G. (2008). Chemistry and the missing era of evolution. Chemistry, 14(13): 3830–3839.Google Scholar
Cairns-Smith, A. G. and Walker, G. L. (1974). Primitive metabolism. Curr. Mod. Biol., 5(4): 173–186.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, 161–180.Google 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, 645–653.Google Scholar
Callaway, E. (2014). Scientists Create First Living Organism with “Artificial” DNA. Nature News, Huffington Post.
Cameron, D. E., Caleb, J., Bashor, J., and Collins, J. (2014). A brief history of synthetic biology. Nature Reviews Microbiology, 12, 381–390.Google Scholar
Capra, F. (2002). The Hidden Connections. Harper Collins.
Capra, F. and Luisi, P. L. (2014). The Systems View of Life: A Unifying Vision. Cambridge University Press.
Caretta, N. (2005). Ipotesi sull'origine della vita: la chimica proteica prebiotica basata su un insieme ridotto di amminoacidi. Thesis 2004/2005, Department of Biology, University Roma Tre.
Carey, M. V. and Small, D. M. (1972). Micelle formation by bile salts. Physical-chemical and thermodynamic considerations. Arch. Intern. Med., 130, 506–527.Google Scholar
Carr, B. (2001). Life, the cosmos and everything. Phys. World, 14, 23–25.Google Scholar
Carr, B., ed. (2007). Universe or Multiverse? Cambridge University Press.
Carrara, P., Stano, P., and Luisi, P. L. (2012). Giant vesicle “colonies”: a model for primitive cell communities. Chembiochem, 13, 1497–1502.Google Scholar
Carrera, J., Elena, S. F., and Jaramillo, A. (2012). Computational design of genomic transcriptional networks with adaptation to varying environments. Proc. Natl. Acad. Sci. USA, 109, 15277–15282.Google Scholar
Caschera, F. and Noireaux, V. (2014) Integration of biological parts toward the synthesis of a minimal cell. Current Opinion in Chemical Biology, 22, 85–91.Google Scholar
Caschera, F., Stano, P., and Luisi, P. L. (2010). Reactivity and fusion between cationic vesicles and fatty acid anionic vesicles. J. Colloid Interface Sci, 345, 561–565.Google Scholar
Caschera, F., Sunami, T., Matsuura, T., et al. (2011). Programmed vesicle fusion triggers gene expression. Langmuir, 27, 13082–13090.Google 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, 102–106.Google Scholar
Cech, T. R. (2011). The RNA worlds in context. Cold Spring Harb Perspect Biol. doi:10.1101/cshperspect.a006742.
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, 1016–1018.Google Scholar
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, 325–329.Google Scholar
Cevc, G. and Blume, G. (1992). Lipid vesicles penetrate into intact skin owing to the transdermal osmotic gradients and hydration force. Biochim Biophys Acta., 1104(1): 226–232.Google Scholar
Cevc, G. and Marsh, D., eds. (1987). Phospholipid Bilayers – Physical Principles and Models, Vol. 5. New York: John Wiley & Sons.
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, 555–559.Google Scholar
Chalmers, D. (1995). Facing up to the problem of consciousness. Journal of Consciousness Studies, 2(3): 200–219.Google Scholar
Chan, L. Y., Kosuri, S., and Endy, D. (September 13, 2005). Refactoring bacteriophage T7. Mol. Syst. Biol., 1: 18.Google Scholar
Chang, M. C. Y. and Keasling, J. D. (2006). Production of isoprenoid pharmaceuticals by engineered microbes. Nature Chemical Biology, 2, 674–681.Google Scholar
Chapman, K. B. and Szostak, J. W. (1995). Isolation of a ribozyme with 5’-5’ ligase activity. Chem. Biol., 2, 325–433.Google Scholar
Chen, I. A. and Szostak, J. W. (2004). A kinetic study of the growth of fatty acid vesicles. Bioph. J., 87, 988–998.Google Scholar
Chen, I. A., Roberts, R. W., and Szostak, J. W. (2004). The emergence of competition between model protocells. Science, 305, 1474–1476.Google Scholar
Chen, I. A., Salehi-Ashtiani, K., and Szostak, J. W. (2005). RNA catalysis in model protocell vesicles. J. Am. Chem. Soc., 127(38): 13213–13219.Google Scholar
Chen, Y., Ma, P., and Gui, S. (2014). Cubic and hexagonal liquid crystals as drug delivery systems. Biomed Res Int. doi:10.1155/2014/815981.
Cheng, J. (2012). Synthesis of polypeptides by ring-opening polymerization of α-aminoacids. Top Curr. Chem., 310: 1–26.Google Scholar
Cheng, Z. and Luisi, P. L. (2003). Coexistence and mutual competition of vesicles with different size distributions. J. Phys. Chem. B, 107(39): 10940–10945.Google Scholar
Chessari, S., Thomas, R., Polticelli, F., and Luisi, P. L. (2006). The production of de novo folded proteins by a stepwise chain elongation: a model for prebiotic chemical evolution of macromolecular sequences. Chemistry & Biodiversity, 3(11): 1202–1210.Google Scholar
Chiarabelli, C. and Luisi, P. L. (2014). Chemical synthetic biology. Science Progress, 97, 48–61.Google Scholar
Chiarabelli, C., Vrijbloed, J. W., Thomas, R. M., and Luisi, P. L. (2006a). Investigation of de novo totally random biosequences, Part I: a general method for in vitro selection of folded domains from a random polypeptide library displayed on phage. Chemistry and Biodiversity, 3, 827–839.Google Scholar
Chiarabelli, C., Vrijbloed, J. W., De Lucrezia, D., et al. (2006b). Investigation of de novo totally random biosequences, Part II: on the folding frequency in a totally random library of de novo proteins obtained by phage display. Chemistry and Biodiversity, 3, 840–859.Google Scholar
Chiarabelli, C., Stano, P., and Luisi, P. L. (2009). Chemical approaches to synthetic biology. Curr. Opin. Biotech., 20, 492–497.Google Scholar
Chiarabelli, C., Stano, P., Anella, F., Carrara, P., and Luisi, P. L. (2012). Approaches to chemical synthetic biology. FEBS Letters, 586, 2138–2145.Google Scholar
Chiarabelli, C., Stano, P., and Luisi, P. L. (2013). Chemical synthetic biology: a mini-review. Frontiers in Microbiotechnology, Ecotoxicology and Bioremediation, 4, 285. doi:10.3389/fmicb.2013.00285.Google Scholar
Christidis, T. (2002). Probabilistic causality and irreversibility: Heraclitus and Prigogine. In Atmanspacher, H. and Bishop, R., eds., Between Chance and Choice. Academic Imprint.
Chungcharoenwattana, S. and Ueno, M. (2004). Size control of mixed egg yolk phosphatidylcholine (EggPC)/oleate. Chem. Pharm. Bull., 52, 1058–1062.Google Scholar
Chungcharoenwattana, S. and Ueno, M. (2005a). New vesicle formation upon oleate addition to preformed vesicles. Chem. Pharm. Bull., 53, 260–262.Google Scholar
Chungcharoenwattana, S. and Ueno, M. (2005b). Effect of preformed egg phosphatidylcholine vesicles on spontaneous vesiculation of oleate micelles. Colloid Pol. Sci., 283, 1180–1189.Google Scholar
Chopra, P. and Kamma, A. (2006). Engineering life through Synthetic Biology. In Silico Biol., 6(5): 401–410.Google Scholar
Church, G. M., Elowitz, M. B., Smolke, C. D., Voigt, C. A., and Weiss, R. (2014). Realizing the potential of synthetic biology. Nature Reviews Molecular Cell Biology, 15, 289–294.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, 125–132.Google Scholar
Chyba, F. and McDonald, G. D. (1995). The origin of life in the solar system: current issues. Ann. Rev. Earth Planet. Sci., 23, 215–249.Google Scholar
Cisar, J. O., Xu, D. Q., Thompson, J., Swaim, W., Hu, L., Kopecko, D. J. (2000). An alternative interpretation of nanobacteria-induced biomineralization. Proc Natl Acad Sci USA, 97, 11511–11515.Google Scholar
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, 1881–1888.Google Scholar
Cohlberg, J. A. and Nomura, M. (1976). Reconstitution of Bacillus stearothermophilus 50S ribosomal subunits from purified molecular components. The Journal of Biological Chemistry, 251, 209–221.Google Scholar
Coleman, P. (2007). Frontier at your fingertips. Nature, 446, 379–385.Google Scholar
Collet, H., Bied, C., Mion, L., and Commeyras, A. (2010). Chem Inform Abstract: A New Simple and Quantitative Synthesis of α-Amino Acid-N- carboxyanhydrides (Oxazolidine-2,5-diones). Chem. Inform., 28(15).Google Scholar
Commeyras, A., Collet, H., Boiteau, L., et al. (2002). Prebiotic synthesis of sequential peptides on the Hadean Beach by a molecular engine working with nitrogen oxides as energy sources. Polymer International, 51, 661–665.Google Scholar
Commeyras, A., Boiteau, L., Vandenabeele-Trambouze, O., and Selsis, F. (2005). Peptide emergence, evolution and selection on the primitive Earth. In Gargaud, M., Barbier, B., Martin, H. and Reisse, J., eds., Lectures in Astrobiology – Vol. I: From Prebiotic Chemistry to the Origins of Life on Earth. Springer-Verlag (Part II, Chap. 4), pp. 517–545.
Conway-Morris, S. (2003). Life's Solution, Inevitable Humans in a Lonely Universe. Cambridge University Press.
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, 4109–4115.Google Scholar
Cooper, G., Kimmich, N., Belisle, W., Sarinana, J., Brabham, K., and Garrel, L. (2001). Carbonaceous meteorites as a source of sugar-related organic compounds for the early Earth. Nature, 414, 879–883.Google Scholar
Cooper, S. J. (2008). From Claude Bernard to Walter Cannon. Emergence of the concept of homeostasis. Appetite, 51(3): 419–427.Google Scholar
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
Coveney, P. and Highfield, R. (1990). The Arrow of Time. W. H. Allen.
Crans, D. C. and Levinger, N. E. (2012). The Conundrum of pH in water nanodroplets: sensing pH in reverse micelle water pools. Acc. Chem. Res., 45, 1637–1645.Google Scholar
Crick, F. (1966). Of Molecules and Men. University of Washington Press.
Crick, F. H. C. (1968). The origin of the genetic code. J Mol Biol., 38, 367–379.Google Scholar
Crick, F. H. C. (1980). The Astonishing Hypothesis. The Search of the Soul from a Chemical Perspective. Scribner.
Cronin, J. R. and Pizzarello, S. (1997). Enantiomeric excesses in meteoritic amino acids. Science, 275, 951–955.Google Scholar
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, 1588–1589.Google Scholar
Cullis, P. R., Hope, M. J., Bally, M. B., et al. (1987). Liposomes as pharmaceuticals. In Ostro, M. J. ed., Liposomes. From Biophysics to Therapeutics. Marcel Dekker, pp. 39–72.
D'Aguanno, E., Altamura, E., Mavelli, F., et al. (2015). Physical routes to primitive cells: An experimental model based on the spontaneous entrapment of enzymes inside micrometer-sized liposomes. MDPI Life, 5, 969–996.Google Scholar
Damasio, A. R. (1999). The Feeling of What Happens. Harcourt.
Damer, B. and Deamer, D. (2015). Coupled phases and combinatorial selection in fluctuating hydrothermal pools: a scenario to guide experimental approaches to the origin of cellular life. Life, 5, 872–887.Google Scholar
Damiano, L. (2006). L'unità in Dialogo: Autoorganizzazione, Autopoiesi, Enazione e Relazione Cognitiva. Doctoral Thesis, University of Bergamo Press.
Damiano, L. and Luisi, P. L. (2010). Verso una ridefinizione autopoietica della vita. Orig. Vita Evol. Biosph., 40, 145–149.Google Scholar
Davidson, A. R. and Sauer, R. T. (1994). Folded proteins occur frequently in libraries of random amino acid sequence. Proc. Natl. Acad. Sci. USA, 91, 2146–2150.Google Scholar
Davidson, A. R., Lumb, K. J., and Sauer, R. T. (1995). Cooperatively folded proteins in random sequence libraries. Nature Structural Biology, 2, 856–864.Google Scholar
Davies, B. (1999). Evolution of the genetic code. Progr. Biophys. Mol. Biol., 72, 157–243.Google Scholar
Davies, B. (2002) Molecular evolution before the origin of species, Progr. Biophys. Mol. Biol., 79, 77–133.Google Scholar
Davies, P. (1999). The Fifth Miracle: The Search for the Origin and Meaning of Life. Simon & Schuster.
Davies, P. (2007). Cosmic Jackpot, Houghton Mifflin; also appeared as The Goldilocks Enigma: Why is the Universe Just Right for Life? Allen Lane, 2006.
Dawkins, R. (1990). The Blind Watchmaker: Why the Evidence of Evolution Reveals a Universe without Design. Penguin Books.
Dawkins, R. (2002). How Life Began: The Genesis of Life on Earth. Cambridge, MA: Foundation for New Directions.
Day, W. (2002). How Life Began: the Genesis of Life on Earth. Cambridge, MA: Foundation for New Directions.
Deamer, D. W. (1985). Boundary structures are formed by organic components of the Murchison carbonaceous chondrite. Nature, 317, 792–794.Google Scholar
Deamer, D. W. (1998). Possible starts for primitive life. In Brack, A., ed., The Molecular Origins of Life. Cambridge University Press.
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.Google Scholar
Deamer, D. W., Harang-Mahon, E., and Bosco, G. (1994). Self-assembly and function of primitive membrane structures. In Bengtson, S., ed., Early Life on Earth. Nobel Symposium No. 84. Columbia University Press, pp. 107–123.
Decher, G. (1997). Fuzzy nano-assemblies: toward layered polymeric multicomposites. Science, 277, 1232–1237.Google 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, 281–291.Google Scholar
de Duve, C. (1991). Blueprint for a Cell: The Nature and the Origin of Life. Neil Patterson Publishers.
de Duve, C. (2002). Life Evolving: Molecules, Mind and Meaning. Oxford University Press.
de Duve, C. (2005). Singularities. Cambridge University Press.
de Duve, C. and Miller, S. (1991). Two-dimensional life? Proc. Natl. Acad. Sci., 88, 10014–10017.Google Scholar
De 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.Google Scholar
De Lucrezia, D., Franchi, M., Chiarabelli, C., Gallori, E., and Luisi, P. L. (2006a). Investigation of de novo totally random biosequences, Part III: RNA Foster: a novel assay to investigate RNA folding structural properties. Chemistry and Biodiversity, 3, 860–868.Google Scholar
De Lucrezia, D., Franchi, M., Chiarabelli, C., Gallori, E., and Luisi, P. L. (2006b). Investigation of de novo totally random biosequences, Part IV: folding properties of de novo, totally random RNAs. Chemistry and Biodiversity, 3, 869–877.Google Scholar
De Napoli, M., Nardis, S., and Paolesse, R. (2004). Hierarchical porphyrin self-assembly in aqueous solution. J. Am. Chem. Soc., 126, 5934–5935.Google Scholar
de Souza, T. P., Stano, P., and Luisi, P. L. (2009). The minimal size of liposome-based model cells brings about a remarkably enhanced entrapment and protein synthesis. Chembiochem Eur. J. Chem. Biol., 10, 1056–1063.Google Scholar
de Souza, T. P., Steiniger, F., Stano, P., Fahr, A., and Luisi, P. L. (2011). Spontaneous crowding of ribosomes and proteins inside vesicles: a possible mechanism for the origin of cell metabolism. Chem. Biochem., 12, 2325–2330.Google Scholar
de Souza, T. P., Stano, P., Steiniger, F., et al. (2012). Encapsulation of ferritin, ribosomes, and ribo-peptidic complexes inside liposomes: insights into the origin of metabolism. Orig. Life Evol. Biospheres, 42, 421–428.Google Scholar
Diedrich, G., Spahn, C. M. T., Stelzl, U., et al. (2000). Ribosomal protein L2 is involved in the association of the ribosomal subunits, trna binding to A and P sites and peptidyl. Embo Journal, 19, 5241–5250.Google Scholar
Diener, T. O. (1971). Potato spindle tuber “virus.” IV. A replicating, low molecular weight RNA. Virology, 45(2): 411–428.Google Scholar
Di Giulio, M. (1998). Reflections on the origin of the genetic code: a hypothesis. J. Theor. Biol., 191, 191–196.Google Scholar
Di Giulio, M. (2001). The non universality of the genetic code: the universal ancestor was a progenote. J. Theor. Biol., 209, 345–349.Google Scholar
Di Giulio, M. (2003). The early phases of the genetic code origin: conjecture on the evolution of coded catalysis. Orig. Life Evol. Biosph., 33, 479–489.Google Scholar
Di Giulio, M. and Medugno, M. (1999). Physicochemical optimization in the genetic code origin as the number of codified amino acids increase. J. Mol. Evol., 49, 1–10.Google Scholar
Dodevski, I., Nucci, N. V., Valentine, K. G., et al. (2014). Optimized reverse micelle surfactant system for high-resolution NMR spectroscopy of encapsulated proteins and nucleic acids dissolved in low viscosity fluids. J. Am. Chem. Soc., 136, 3465–3474.Google Scholar
Doi, N., Kakukawa, K., and Yanagawa, H. (2005). High solubility of random-sequence proteins consisting of five kinds of primitive amino acids. Protein Engineering, Design & Selection, 18, 279–84.Google Scholar
Dolgin, E. (2015). Synthetic biology: Safety boost for GM organisms. Nature, 517 (7535): 423. doi: 10.1038 / 517423a.Google Scholar
Domazou, A. S. and Luisi, P. L. (2002). Size distribution of spontaneously formed liposomes by the alcohol injection method. J. Liposome Res., 12(3): 205–220.Google Scholar
Dominak, L. M., Omiatek, D. M., Gundermann, E. L., Heien, M. L., and Keating, C. D. (2010). Polymeric crowding agents improve passive biomacromolecule encapsulation in lipid vesicles. Langmuir, 26 (16): 13195–131200.Google Scholar
Dubois, L. H. and Nuzzo, R. G. (1992). Synthesis, structure, and properties of model organic-surfaces. Ann. Rev. Phys. Chem., 43, 437–463.Google 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, 815–819.Google Scholar
Dymond, J. S., Richardson, S. M., Coombes, C. E., et al. (2011). Synthetic chromosome arms function in yeast and generate phenotypic diversity by design. Nature, 477 (7365): 471–476.Google Scholar
Dyson, F. J. (1985). Origins of Life. Cambridge University Press.
Eddy, S. R. (2002). Non-coding RNA genes and the modern RNA world. Cell, 109, 137–140.Google 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, 245–251.Google Scholar
Eigen, M. (1971). Self-organization of matter and the evolution of biological macromolecules. Naturwissenschaften, 58, 465–523.Google Scholar
Eigen, M. and Schuster, P. (1977). Hypercycle – principle of natural self-organization. A. Emergence of hypercycle. Naturwissenschaften, 64, 541–565.Google Scholar
Eigen, M. and Schuster, P. (1979). The Hypercycle: A Principle of Natural Self-Organization. Springer Verlag.
Eigen, M. and Winkler-Oswatitisch, R. (1992). Steps Towards Life. Oxford University Press.
Eigen, M., Gardiner, W., Schuster, P., and Winkler-Oswatitsch, R. (1981). The origin of genetic information. Sci. Am., 244(4): 88–92.Google Scholar
El Seoud, O. A. (1984). In Luisi, P. L. and Straub, B., eds., Reverse Micelles. Plenum Press.
Ellis, G. (2005). Physics ain't what it used to be. Nature, 438, 739–740.Google Scholar
Engels, F. (1877). Anti-Duehring. Translated by Lewis, Austin. Chicago: Charles H. Kerr & Company, 1907.
Engels, F. (1883). Dialectics of Nature. Notes and Fragments. Translated by Dutt, Clemens. Moscow: Progress Publishers, edn., 1934. Included in Karl Marx and Frederick Engels, Collected Works, Volume 25 (Engels), published in 1987 by Lawrence & Wishart.
Engels, F. (1894). Herrn Eugen Dühring's Umwalzung der Wissenschaft. Dietz Verlag. English translation (Herr Eugen Dühring's Revolution in Science) was included in: Karl Marx and Friedrich Engels, Collected Works. Volume 25 (Engels), published in 1987 by Lawrence & Wishart.
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, 23–27.Google Scholar
Erwin, D. H. (2003). Life's solution – inevitable humans in a lonely universe. Science, 302, 1682–1683.Google Scholar
Eschenmoser, A. (1999). Chemical etiology of nucleic acid structure. Science, 284, 2118–2124.Google Scholar
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.
Eschenmoser, A. and Kisakürek, M. V. (1996). Chemistry and the origin of life. Helv. Chim. Acta., 79, 1249–1259.Google 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, 1277–1282.Google Scholar
Falbe, J. (1987). Surfactants in Consumer Products. Theory, Technology and Applications. Springer Verlag.
Famiglietti, M., Hochköppler, A., Wehrli, E., and Luisi, P. L. (1992). Photosynthetic activity of cyanobacteria in water-in-oil microemulsions. Biotechnol. Bioeng., 40, 173–178.Google Scholar
Famiglietti, M., Hochköppler, A., and Luisi, P. L. (1993). Surfactant-induced hydrogen production in cyanobacteria. Biotechnol. Bioeng., 42, 1014–1018.Google Scholar
Fan, K. and Wang, W. (2003). What is the minimum number of letters required to fold a protein? J. Mol. Biol., 328, 921–926.Google Scholar
Fanelli, D. and McKane, A. J. (2008). Thermodynamics of vesicle growth and instability. Physical Review E, 78, 051406.Google Scholar
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.
Ferris, J. P. (1998). Catalyzed RNA synthesis for the RNA world. In Brack, A., ed., The Molecular Origin of Life. Cambridge University Press, pp. 255–256.
Ferris, J. P. and Ertem, G. (1992). Oligomerization reaction of ribonucleosides on montmorillonite: reaction of 5′-phosphorimidazolide of adenosine. Science, 257, 1387–1389.Google 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, 12270–12275.Google Scholar
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.Google Scholar
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, 511–518.Google Scholar
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, 225–231.Google 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.Google Scholar
Field, R. J. (1972). A reaction periodic in time and space. J. Chem. Educ., 49, 308–311.Google Scholar
Fikes, B. J. (2014). Life engineered with expanded genetic code. The San Diego Union-Tribune.
Fiordemondo, D. and Stano, P. (2007). Lecithin-based water-in-oil compartments as dividing bioreactors. ChemBioChem, 8, 1965–1973.Google Scholar
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, 177–188.Google Scholar
Fleischaker, G. (1988). Autopoiesis: the status of its system logic. Biosystems, 22, 37–49.Google Scholar
Fletcher, P. D. and Robinson, B. H. (1981). Ber. Bunsenges. Phys. Chem., 85, 863.
Foldvari, M., Geszles, A., and Mezei, M. (1990). J. Microencapsul., 7, 479–489.
Folk, R.L. (1993). Sem imaging of bacteria and nannobacteria in carbonate sediments and rocks. Journal of sedimentary petrology, 63, 990–999.Google Scholar
Folsome, C. E. (1979). The Origin of Life: A Warm Little Pond. W. H. Freeman & Co.
Fontell, K. (1990). Cubic phases in surfactant and surfactant-like lipid systems. Colloid Polym. Sci., 268, 265–285.Google Scholar
Forster, A. C., and Symons, R. H. (1987). Self-cleavage of plus and minus RNAs of a virusoid and a structural model for the active sites. Cell, 49(2): 211–220.Google Scholar
Forster, C. A., and Church, G. M. (2006). Towards synthesis of a minimal cell. Mol. Syst. Biol., 2, 45. doi: 10.1038/msb4100090.Google Scholar
Forster, C. A., and Church, G. M. (2007). Synthetic biology projects in vitro. Genome Research, 17, 1–17.Google Scholar
Föster, S. and Plantenberg, T. (2002). From self-organizing polymers to nanohybrid and biomaterials. Angew. Chem. Int. Ed. Engl., 41, 688–714.Google Scholar
Fox, G. E., Tran, Q., and Yonath, A. (2012). An exit cavity was crucial to the polymerase activity of the early ribosome. Astrobiology, 12, 57–60.Google Scholar
Fox, S. W. (1988). The Emergence of Life. Basic Books.
Fox, S. W. and Dose, K. (1972). Molecular Evolution and the Origin of Life. W. H. Freeman.
Fox-Keller, E. (2002). The Century of the Gene. Harvard University Press.
Fraenkel-Conrat, H. and Williams, R. C. (1955). Reconstitution of active tobacco mosaic virus from its inactive protein and nucleic acid components. Proc. Nat. Acad. Sci. USA., 41, 690–698.Google Scholar
Franceschi, F. J. and Nierhaus, K. H. (1990). Ribosomal protein-l15 and protein-l16 are mere late assembly proteins of the large ribosomal-subunit – analysis of an Escherichia coli mutant lacking l15. Journal of Biological Chemistry. 265, 16676–16682.Google Scholar
Franz, M.-L. von (1988). Psyche und Materie. Daimon Verlag.
Fraser, C. M., Gocayne, J. D., White, O., et al. (1995). The minimal gene complement of Mycoplasma genitalium. Science, 270, 397–403.Google Scholar
Freitas, R. A. Jr., and Merkle, R. C. (2004). Kinematic Self-Replicating Machines. Landes Bioscience.
Fry, I. (1999). The Emergence of Life on Earth: A Historical and Scientific Overview. London: Free Association Books.
Fry, I. (2000). Emergence of Life on Earth: A Historical and Scientific Overview. New Brunswick, NJ: Rutgers University Press.
Fry, I. (2011). The role of natural selection in the origin of life. Origins of Life and Evolution of Biospheres, 41(1): 3–16.Google Scholar
Fujii, S., Matsuura, T., Sunami, T., Kazuta, Y., and Yomo, T. (2015). In vitro directed evolution of alpha-hemolysin by liposome display. Biophysics, 11: 67–72.Google Scholar
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, 439–468.Google Scholar
Ganti, T. (1975). Organization of chemical reactions into dividing and metabolizing units: the chemotons. BioSystems, 7, 15–21.Google Scholar
Ganti, T. (1984). Chemoton elmélet 1. kötet. A fluid automaták elméleti alapjai. Translated as Chemoton Theory, Vol. 1., Theory of Fluid Automata. OMIKK.
Ganti, T. (2003). The Principles of Life. Oxford University Press.
Gao, X. and Huang, L. (1995). Cationic liposome-mediated gene transfer. Gene Ther., 2(10): 710–722.Google Scholar
Gardner, P. M. and Davis, B. G. (2011). Approaches to building chemical cells/chells: examples of relevant mechanistic “couples.” In Luisi, P. L. and Stano, P., eds., The Minimal Cell. Springer.
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, 537–552.Google Scholar
Gennis, R. B. (1989). Biomembranes, Molecular Structure and Function. Springer Verlag.
Ghosh, I. and Chmielewski, J. (2004). Peptide self-assembly as a model of proteins in the pre-genomic world. Curr. Opin. Chem. Biol., 8, 640–644.Google Scholar
Gibson, D. G., Benders, G. A., Andrews-Pfannkoch, C., et al. (2008a). Complete chemical synthesis, assembly, and cloning of a Mycoplasma genitalium genome. Science, 319(5867): 1215–20.Google Scholar
Gibson, D. G., Benders, G. A., et al. (2008b). One-step assembly in yeast of 25 overlapping DNA fragments to form a complete synthetic Mycoplasma genitalium genome . Proc. Natl. Acad. Sci. USA., 105(51): 20404–204009. doi:10.1073/pnas.0811011106.Google Scholar
Gibson, D. G., Glass, J. I., Lartigue, C., et al. (2010). Creation of a bacterial cell controlled by a chemically synthesized genome. Science, 329(5987): 52–56. doi:10.1126/science.1190719.Google Scholar
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, 518–537.Google Scholar
Gilbert, R. J. C., Fucini, P., Connell, S., et al. (2004). Three-dimensional structures of translating ribosomes by cryo-EM. Molecular Cell. 14, 57–66.Google Scholar
Gilbert, W. (1986). The RNA world. Nature, 319, 618.Google Scholar
Gold, T. (1979). Terrestrial sources of carbon and earthquake outgassing. Journal of Petroleum Geology 1(3): 3–19.Google Scholar
Glotzer, S. C. (2004). Materials science. Some assembly required. Science, 306, 419–420.Google Scholar
Goodenough, U. and Deacon, T. W. (2006). Emergence and religious naturalism. In Clayton, P., ed., Oxford Handbook of Science and Religion. Oxford University Press.
Gorlero, M., Wieczorek, R., Adamala, K., et al. (2009). Ser-His catalyses the formation of peptides and PNAs. FEBS Letters, 583, 153–156.Google Scholar
Gould, S. J. (1989). Wonderful Life. Penguin Books.
Graf, A., Winterhalter, M., and Meier, W. (2001). Nanoreactors from polymer-stabilized liposomes. Langmuir, 17, 919–923.Google Scholar
Green, R. and Noller, H.F. (1997). Ribosomes and translation. Annual Review of Biochemistry, 66, 679–716.Google Scholar
Gregoriadis, G. (1976a). The carrier potential of liposomes in biology and medicine (first of two parts). New Engl. J. Med., 295, 704–710.Google Scholar
Gregoriadis, G. (1976b). The carrier potential of liposomes in biology and medicine (second of two parts). New Engl. J. Med., 295, 765–770.Google Scholar
Gregoriadis, G., ed. (1988). Liposomes and Carriers of Drugs: Recent Trends and Progress. New York: John Wiley & Sons.
Gregoriadis, G. (1995). Engineering liposomes for drug delivery: progress and problems. Trends Biotechnol., 13(12): 527–537.Google Scholar
Groen, J., Deamer, D. W., Kros, A., and Ehrenfreund, P. (2012). Polycyclic aromatic hydrocarbons as plausible prebiotic membrane components. Orig. Life Evol. Biosph., 42(4): 295–306.Google Scholar
Habraken, G. J. M., Peeters, M., Dietz, C. H. J. T., Koninga, C. E., and Heise, A. (2010). How controlled and versatile is N-carboxy anhydride (NCA) polymerization at 0 °C? Effect of temperature on homo-, block- and graft (co)polymerization. Polym. Chem., 1, 514–524.Google Scholar
Häckel, E. (1866). Allgemeine Anatomie der Organismen. Walter de Gruyer.
Haines, T. H. (1983). Anionic lipid headgroups as a proton-conducting pathway along the surface of membranes: a hypothesis. Proc. Natl. Acad. Sci. USA, 80, 160–164.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
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, 601–606.Google Scholar
Hampl, H., Schulze, H., and Nierhaus, K.H. (1981). Ribosomal components from escherichia-coli 50-s subunits involved in the reconstitution of peptidyltransferase activity. Journal of Biological Chemistry, 256, 2284–2288.Google Scholar
Han, D. and Rhee, J. S. (1986). Biotechnol. Bioeng., 27, 1250–1255.
Hanczyc, M. M. and Szostak, J. W. (2004). Replicating vesicles as models of primitive cell growth and division. Curr. Opin. Chem. Biol., 8(6): 660–664.Google Scholar
Hanczyc, M. M., Fujikawa, S. M., and Szostak, J. W. (2003). Experimental models of primitive cellular compartments: encapsulation, growth, and division. Science, 302, 618–622.Google Scholar
Hansen, J., Mailand, E., Swaminathan, K. K., Schreiber, J., Angelici, B., and Benenson, Y. (2014). Transplantation of prokaryotic two-component signaling pathways into mammalian cells. Proc. Natl. Acad. Sci. USA, 111(44): 15705–15710. doi:10.1073/pnas.1406482111.Google Scholar
Hansler, M. and Jakubke, H.-D. (1996). Nonconventional protease catalysis in frozen aqueous solutions. J. Pept. Sci., 2, 279–289.Google Scholar
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, 2098–2102.Google Scholar
Hargreaves, W. R. and Deamer, D. W. (1978a). Liposomes from ionic, single-chain amphiphiles. Biochemistry, 17, 3759–3768.Google Scholar
Hargreaves, W. R. and Deamer, D. W. (1978b). In Deamer, D. W., ed., Light Transducing Membranes: Structure, Function and Evolution. Academic Press, pp. 23–59.
Hargreaves, W. R., Mulvhill, S. J., and Deamer, D. W. (1977). Synthesis of phospholipids and membranes in prebiotic conditions. Nature, 266, 78–80.Google 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, 911–915.Google Scholar
Häring, G., Pessina, A., Meussdoerffer, F., Hochköppler, 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
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, 7638–7647.Google Scholar
Hayatsu, R., Studier, M. H., Moore, L. P., and Anders, E. (1975). Purines and triazines in the Murchison meteorite. Geochim. Cosmochim. Acta, 39, 471–488.Google Scholar
Hecht, M. H., Das, A., Go, A., Bradley, L. H., and Wei, Y. (2004). De novo proteins from designed combinatorial libraries. Protein Science, 13, 1711–1723.Google 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
Herz-Fischler, R. (1998). A Mathematical History of the Golden Number. New York: Dover.
Hilborn, R. C. (1994). Chaos and Non Linear Dynamics. Oxford University Press.
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, 459–466.Google Scholar
Hirwschmann, H. and Hanson, K. R. (1971). Top Stereochem., 36: 329–399.
Hochköppler, A. and Luisi, P. L. (1989). Solubilization of soybean mitochondria in AOT/isooctane water-in-oil microemulsions. Biotechnol. Bioeng., 33, 1477–1481.Google Scholar
Hochköppler, A. and Luisi, P. L. (1991). Photosynthetic activity of plant cells solubilized in water-in-oil microemulsions. Biotechnol. Bioeng., 37, 918–921.Google 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, 996–997.Google Scholar
Holland, J. H. (1998). Emergence: From Chaos to Order. Oxford University Press.
Holm, N. G. and Andersson, E. M. (1998). Hydrothermal systems. In Brack, A., ed., The Molecular Origin of Life. Cambridge University Press.
Horowitz, N. and Miller, S. (1962). Origins of life: the primal self-organization. In Zechmeister, L., ed., Progress in the Chemistry of Natural Products, Vol. 20. Springer Verlag, pp. 423–459.
Horowitz, P. and Sagan, C. (1993). Five years of Project META: an all-sky narrow-band radio search for extraterrestrial signals. Astrophys. J., 415, 218–233.Google Scholar
Hosoda, K., Sunami, T., Kazuta, Y., Matsuura, T., Suzuki, H., and Yomo, T. (2008). Quantitative study of the structure of multilamellar giant liposomes as a container of protein synthesis reaction. Langmuir, 24, 13540–13548.Google Scholar
Hoyle, F. and Wickramasinghe, C. (1999). Astronomical origins of life – steps towards panspermia. Astrophys. Space Sci., 268, Preface, VII–VIII.Google Scholar
Hoyle, F. and Wickramasinghe, C. (2000). Astronomical Origins of Life – Steps Towards Panspermia. Dordrecht, NL: Kluwer Academic.
Huang, S. S. (1959). Occurrence of life in the universe. Amer. Sci., 47, 397–402.Google Scholar
Huber, C. and Wächtershäuser, G. (1997). Activated acetic acid by carbon fixation on (Fe, Ni)S under primordial condition. Science, 276, 245–247.Google Scholar
Huntley, H. E. (1970). The Divine Proportion: A Study in Mathematical Beauty. New York: Dover.
Hutchinson, C. A., Peterson, S. N., Gill, S. R., et al. (1999). Global transposon mutagenesis and a minimal mycoplasma genome. Science, 286, 2165–2169.Google Scholar
Ikehara, K. (2002). Origins of gene, genetic code, protein and life: comprehensive view of life systems from a GNC-SNS primitive genetic code hypothesis; J. Biosci. 27, 165–186.Google Scholar
Ikehara, K. (2005). Possible steps to the emergence of life: the (GADV)-protein world hypothesis. The Chemical Record, 5, 107–118.Google Scholar
Ikehara, K. (2009). Pseudo-replication of [GADV]-proteins and origin of life. Int. J. Mol. Sci., 10, 1525–1537.Google Scholar
Ikehara, K., Omori, Y., Arai, R., and Hirose, A. (2002). A novel theory on the origin of the genetic code: a GNC-SNS hypothesis. J. Mol. Evol., 54, 530–538.Google Scholar
Imre, V. E. and Luisi, P. L. (1982). Solubilization and condensed packaging of nucleic acids in reversed micelles. Biochem. Biophys. Res. Commun., 107, 538–545.Google Scholar
Ishikawa, K., Sato, K., Shima, Y., Urabe, I., and Yomo, T. (2004). Expression of cascading genetic network within liposomes. FEBS Lett., 576, 387–390.Google Scholar
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): 243–256.Google Scholar
Israelachvili, J. N. (1992). Intermolecular and Surface Forces, edn. Academic Press.
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.Google Scholar
Issac, R. and Chmielewski, J. (2002). Approaching exponential growth with a self-replicating peptide. J. Am. Chem. Soc., 124, 6808–6809.Google Scholar
Itaya, M. (1995). An estimation of the minimal genome size required for life. FEBS Lett., 362, 257–260.Google Scholar
Itojima, Y., Ogawa, Y., Tsuno, K., Handa, N., and Yanagawa, H. (1992). Spontaneous formation of helical structures from phospholipid-nucleoside conjugates. Biochemistry, 31, 4757–4765.Google Scholar
Jacob, F. (1982). The Possible and the Actual. University of Washington Press.
Jaeger, 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, 14712–14717.Google Scholar
Jakubke, H.-D. (1987). Peptides: design, synthesis, and biological activity. In Udenfried, S. and Meienhofer, J., eds., The Peptides: Analysis, Synthesis, Biology, Vol. 9. Academic Press.
Jakubke, H.-D. (1995). Hydrolysis and formation of peptides. In Drauz, K. and Waldmann, H., eds., Enzyme Catalysis in Organic Synthesis, Vol. 1. Wiley-VCH, pp. 431–458.
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.Google 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, 455–464.Google Scholar
Janiak, M. J., Small, D. M., and Shipley, G. G. (1976). Nature of the thermal pretransition of synthetic phospholipids: dimyristoyl- and dipalmitoyllecithin. Biochemistry, 15, 4575–4580.Google Scholar
Jeon, K. W., Lorch, I. J., and Danielli, J. F. (1970). Reassembly of living cells from dissociated components. Science, 167(3925): 1626–1627.Google Scholar
Jewett, M. C., Calhoun, K. A., Voloshin, A., Wuu, J. J., and Swartz, J. R. (2008). An integrated cell-free metabolic platform for protein production and synthetic biology. Mol. Syst. Biol., 4, 220.Google Scholar
Jimenez-Prieto, R., Silva, M., and Perez-Bendito, D. (1998). Approaching the use of oscillating reactions for analytical monitoring. Analyst, 123, 1R–8R.Google Scholar
Jiménez, J. I., Xulvi-Brunet, R., Campbell, G. W., Turk-MacLeod, R., and Chen, I. A. (2013). Comprehensive experimental fitness landscape and evolutionary network for small RNA. Proc. Natl. Acad. Sci. USA, 110, 14984–14991.Google Scholar
Johnson, E. T., and Schmidt-Dannert, C. (2008). Light-energy conversion in engineered microorganisms. Trends Biotechnol., 26(12): 682–689.Google Scholar
Johnston, W. K., Unrau, P. J., Lawrence, M. S., Glasner, M. E., and Bartel, D. P. (2001). RNA-catalyzed RNA polymerization: accurate and general RNA-templated primer extension. Science, 292(5520): 1319–1325.Google 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, 433–441.Google Scholar
Juarrero, A. and Rubino, C. A. (2010). Emergence, Complexity, and Self-Organization: Precursors and Prototypes (Exploring Complexity). Paperback. ISCE Publishing.
Kajander, E. O. and Çiftçioglu, N. (1998). Nanobacteria: An alternative mechanism for pathogenic intra- and extracellular calcification and stone formation. Proc. Natl. Acad. Sci. USA, 95(14): 8274–8279.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, 1371–1374.Google Scholar
Kamtekar, S., Shiffer, J. M., Xiong, H. Y., Babik, J. M., and Hecht, M. H. (1993). Protein design by binary patterning of polar and nonpolar amino acids. Science, 262, 1680–1685.Google Scholar
Kaszuba, M. and Jones, M. N. (1999). Hydrogen peroxide production from reactive liposomes encapsulating enzymes. Biochim. Biophys. Acta, 1419, 221–228.Google Scholar
Kato, A., Yanagisawa, M., Sato, Y. T., Fujiwara, K., and Yoshikawa, K. (2012). Cell-sized confinement in microspheres accelerates the reaction of gene expression. Scientific Report, 2, 283.Google Scholar
Kauffman, S. A. (1986). Autocatalytic set of proteins. J. Theor. Biol., 119, 1–24.Google Scholar
Kauffman, S. A. (1993). The Origins of Order: Self Organization and Selection in Evolution. Oxford University Press.
Kawamura, K. (2002). The origin of life from the life of subjectivity. In Palyi, G., Zucchi, C., and Caglioti, L., eds., Fundamentals of life. Elsevier, pp. 56–76.
Kawamura, K. and Kamoto, F. (2000). Condensation reaction of hexanucleotides containing guanine and cytosine with water soluble carbodiimide. Nucleic Acid Symp. Ser., 44, 217–218.Google Scholar
Kent, S. (1999). Chemical protein synthesis by solid phase ligation of unprotected peptide segments. J. Am. Chem. Soc., 121, 8720–8727.Google Scholar
Kenyon, D. and Mills, G. (1996). The RNA world: a critique. Review Article Origins & Design, 17: 1.Google Scholar
Khalil, A. S. and Collins, J. J. (2010). Synthetic biology: applications come of age. Nature Reviews Genetics, 11, 367–379. doi:10.1038/nrg2775.Google Scholar
Kiedrowski, G. von (1986). A self-replicating hexadeoxynucleotide. Angew. Chem. Int. Ed. Engl., 25, 932–925.Google Scholar
Kiedrowski, G. von (1993). Minimal replicator theory I: parabolic versus exponential growth. In Berlin, D. H., ed., Bioorganic Chemistry, Vol. 3. Springer Verlag, pp. 115–146.
Kikuchi, A., Aoki, Y., Sugaya, S., et al. (1999). Development of novel cationic liposomes for efficient gene transfer into peritoneal disseminated tumor. Human Gene Therapy, 10(6): 947–955.Google Scholar
Kim, J. (1984). Concepts of supervenience. Phil. Phen. Res., 45, 153–176.Google Scholar
Kimura, M. (1983). The Neutral Theory of Molecular Evolution. Cambridge University Press.
Kita, H., Matsuura, T., Sunami, T., Hosoda, K., Ichihashi, N., et al. (2008). Replication of genetic information with self-encoded replicase in liposomes. Chem. Bio. Chem., 9, 2403–2410.Google Scholar
Kitano, H. (2007). Towards a theory of biological robustness. Molecular Systems Biology, 3(1).Google Scholar
Kitney, R. and Freemont, P. (2012). Synthetic biology – the state of play, FEBS Letters, 586, 2029–2036.Google Scholar
Klee, R. (1984). Micro-determinism and concepts of emergence. Phil. Sci., 51, 44–63.Google Scholar
Knenvolden, K., Lawless, J. G., Pering, K., et al. (1970). Evidence for extraterrestrial amino acids and hydrocarbons in the Murchison meteorite. Nature, 228, 923–926.Google Scholar
Kobayashi, K. and Kanaizuka, Y. (1977). Reassembly of living cells from dissociated components in Bryopsis. Plant & Cell Physiol., 18, 1373–1377.Google Scholar
Kobayashi, K., Tsuchiya, M., Oshima, T., and Yanagawa, H. (1990). Abiotic synthesis of amino acids and imidazole by proton irradiation of simulated primitive earth atmospheres. Origins of Life and Evolution of the Biosphere, 20, 99–109.Google Scholar
Kobayashi, K., Kaneko, T., Saito, T., and Oshima, T. (1998). Amino acid formation in gas mixtures by high-energy particle irradiation. Origins of Life and Evolution of the Biosphere, 28, 155–165.Google Scholar
Kolisnychenko, V., Plunkett, G. III, Herring, C. D., et al. (2002). Engineering a reduced Escherichia coli genome. Genome Res., 12, 640–647.Google Scholar
Kondepudi, D. K. and Prigogine, I. (1981). Sensitivity of non-equilibrium systems. Physica A, 107, 1–24.Google Scholar
Kondepudi, D. K., Prigogine, I., and Nelson, G. (1985). Sensitivity of branch selection in nonequilibrium systems. Phys. Lett. A, 111, 29–32.Google Scholar
Kondepudi, D. K., Kaufman, R., and Singh, N. (1990). Chiral symmetry breaking in sodium chlorate crystallization. Science, 250, 975.Google 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, 2380–2384.Google Scholar
Kool, E. T. and Morales, J. (2005). Efficient replication between non-hydrogen-bonded nucleoside shape analogs. In Simon, M., ed., Emergent Computation: Emphasizing Bioinformatics. Springer, pp. 88–98.
Koonin, E. V. (2000). How many genes can make a cell: the minimal-gene-set concept. Annu. Rev. Genomics Human Genet., 1, 99–116.Google Scholar
Koshland, D. E. Jr. (2002). The seven pillars of life. Science, 295, 2215–2216.Google Scholar
Krupkin, M., Bashan, A., and Yonath, A. (2014). Glimpse into the origin of life: what was first, the genetic code or its products, the proteins? In Trueba, G., ed., Why Does Evolution Matter? The Importance of Understanding Evolution. Cambridge Scholars Publishing, pp. 87–100.
Kuiper, T. B. H. and Morris, M. (1977). Searching for extraterrestial civilizations. Science, 196, 616–621.Google Scholar
Kullmann, W. (1987). Enzymatic Peptide Synthesis. CRC Press.
Kunin, V. (2000). A system of two polymerases – a model for the origin of life. Origins of Life and Evolution of the Biosphere, 30(5): 459–468.Google Scholar
Kuruma, Y., Stano, P., Ueda, T., and Luisi, P. L. (2009). A synthetic biology approach to the construction of membrane proteins in semi-synthetic minimal cells. Biochim. Biophys. Acta, 1788, 567–574.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, 2533–2536.Google Scholar
Lancaster, W. A. and Adams, M. W. W. (2011). The influence of environment and metabolic capacity on the size of a microorganism. In Luisi, P. L. and Stano, P., eds., The Minimal Cell. Netherlands: Springer, pp. 93–103.
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, 2102–2106.Google 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, 173–178.Google Scholar
Lane, N. (2009). Power, Sex, Suicide. Oxford University Press.
Langton, C. G., ed. (1988). Artificial Life: Proceedings of an Interdisciplinary Workshop on the Synthesis and Simulation of Living Systems. Addison-Wesley.
Langton, C. G. (1990). Computation at the edge of chaos: phase transitions and emergent computation. Physica, D42, 12–37.Google Scholar
Langton, C. G., ed. (1993). Artificial Life III: Proceedings of the Third Interdisciplinary Workshop on the Synthesis and Simulation of Living Systems. Addison-Wesley.
Langton, C. G., ed. (1995). Artificial Life: An Overview. MIT Press.
Lapique, N. and Benenson, Y. (2014). Digital switching in a biosensor circuit via programmable timing of gene availability. Nature Chemical Biology. doi:10.1038/nchembio.1680.
Larsson, K. (1989). Cubic lipid-water phases: structures and biomembrane aspects. J. Phys. Chem., 93, 7304–7314.Google Scholar
Lartigue, C., Glass, J. I., Alperovich, N., et al. (2007). Genome transplantation in bacteria: changing one species to another. Science, 317(5838): 632–638.Google Scholar
Lasch, J., Laub, R., and Wohlrab, W. (1991). How deep do intact liposomes penetrate into human skin? J. Controll. Release, 18, 55–58.Google Scholar
Lasic, D. D. (1995). In Lipowsky, R. and Sackmann, E., eds., Handbook of Biological Physics, Vol. 1. Elsevier, pp. 491–519.
Lawless, J. G. and Yuen, G. U. (1979). Quantification of monocarboxylic acids in the Murchison carbonaceous meteorite. Nature, 282, 396–398.Google Scholar
Lawrence, D. S., Jiang, T., and Levett, M. (1995). Self-assembling supramolecular complexes. Chem. Rev., 95, 2229–2260.Google Scholar
Lazcano, A. (2003). Just how pregnant is the universe? Science, 299, 347–348.Google Scholar
Lazcano, A. (2004). An answer in search of a question. A review of How Life Began: The Genesis of Life on Earth, by William Day. Astrobiology, 4(4): 469–471.Google Scholar
Lazcano, A., and Bada, J. L. (2003). The 1953 Stanley L. Miller experiment: fifty years of prebiotic organic chemistry. Orig. Life Evol. Biosph., 33, 235–242.Google Scholar
Lazzara, S. (2001). Vedi Alla Voce Scienza. Manifesto Libri.
le Doux, J. (2002). Synaptic Self: How Our Brains Become Who We Are. Viking Books.
Lee, D. H., Granja, J. R., Martinez, J. A., Severin, K., and Ghadiri, M. R. (1996). A self-replicating peptide. Nature, 382, 525–528.Google Scholar
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, 591–594.Google Scholar
Lee, S. K., Chou, H., Ham, T. S., Lee, T. S., and Keasling, J. D. (2008). Metabolic engineering of microorganisms for biofuels production: from bugs to synthetic biology to fuels. Curr Opin Biotechnol., 19(6):556–563.Google Scholar
Lehn, J.-M. (2002). Toward self-organization and complex matter. Science, 295(5564): 2400–2403. doi:10.1126/science.1071063.Google Scholar
Leonard, E., Nielsen, D., Solomon, K., and Prather, K. J. (2008). Engineering microbes with synthetic biology frameworks. Trends Biotechnol., 12, 674–681.Google Scholar
Leser, M. E. and Luisi, P. L. (1989). Liquid 3-phase micellar extraction of peptides. Biotech. Techniques, 3, 149–154.Google Scholar
Leser, M. E. and Luisi, P. L. (1990). Application of reverse micelles for the extraction of amino acids and proteins. Chimia, 44, 270–282.Google Scholar
Levashov, A. V., Klyachko, N. L., Psbezhetski, A. V., et al. (1989). Biochim. Biophys. Acta, 988, 221–256.Google Scholar
Levy, M. and Ellington, A. D. (2003). Peptide-template nucleic acid ligation. J. Mol. Evol., 56, 607–615.Google Scholar
Lewontin, R. C. (1970). The units of selection. Annu. Rev. Ecol. Syst., 1, 1–18.Google Scholar
Lewontin, R. C. (1993). The Doctrine of DNA – Biology as an Ideology. Penguin Books.
Li, J., Ballmer, S. G., Gillis, E. P., et al. (2015). Synthesis of many different types of organic small molecules using one automated process, Science, 347(6227): 1221–1226.Google Scholar
Li, T. and Nicolaou, K. C. (1994). Chemical self-replication of palindromic duplex DNA. Nature, 369, 218–221.Google Scholar
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, 2675–2680.Google Scholar
Li, X. and Chmielewski, J. (2003). Peptide self-replication enhanced by a proline kink. J. Am. Chem. Soc., 125, 11820–11821.Google Scholar
Lifson, S. (1997). On the crucial stages in the origin of animate matter. J. Mol. Evol., 44, 1–8.Google Scholar
Lindahl, T. (1993). Instability and decay of the primary structure of DNA. Nature, 362(6422): 709–715.Google Scholar
Lindblom, G. and Rilfors, L. (1989). Cubic phases and isotropic structures formed by membrane lipids – possible biological relevance. Biochem. Biophys. Acta, 988, 221–256.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, 153–180.Google Scholar
Liu, A. P. and Fletcher, D. A. (2009). Biology under construction: in vitro reconstitution of cellular function. Nature Reviews, 10, 644–670.Google Scholar
Livio, M. (2002). La Sezione Aurea. Storia di un numero e di un mistero che dura da tremila anni, Rizzoli.
Lohrmann, R. and Orgel, L. E. (1973). Prebiotic activation processes. Nature, 244, 418.Google 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, 10910–10916.Google Scholar
Longo, L. M., Lee, J., and Blaber, M. (2013). Simplified protein design biased for prebiotic amino acids yields a foldable, halophilic protein. PNAS, 110(6): 2135–2139.Google Scholar
Love, S. G. and Brownlee, D. E. (1993). A direct measurement of the terrestrial mass accretion rate of cosmic dust. Science, 262, 550–553.Google Scholar
Lovelock, J. E. (1979). Gaia: A New Look at Life on Earth. Oxford University Press.
Luci, P. (2003). Gene cloning expression and purification of membrane proteins. ETH-Z Dissertation Nr. 15108, Zurich.
Lucks, J. B., Qi, L., Whitaker, W. R., and Arkin, A. P. (2008). Toward scalable parts families for predictable design of biological circuits. Curr. Opin. Microbiol., 11(6): 567–573.Google Scholar
Luhmann, K. (1984). Soziale Systeme. Suhrkamp.
Luisi, P. L. (1979). Why are enzymes macromolecules? Naturwissenschaften, 66, 498–504.Google Scholar
Luisi, P. L. (1985). Enzyme hosted in reverse micelles in hydrocarbon solution. Angew. Chem., 24, 439–450.Google Scholar
Luisi, P. L. (1993). Defining the transition to life: self-replicating bounded structures and chemical autopoiesis. In Stein, W. and Varela, F. J., ed., Thinking About Biology, SFI Studies in the Sciences of Complexity. Addison-Wesley-Longman.
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, 425–438.Google Scholar
Luisi, P. L. (1997). Self-reproduction of chemical structures and the question of the transition to life. In Cosmovici, C. B., Bowyer, S., and Werthimer, D., eds., Astronomical and Biochemical Origins and the Search for Life in the Universe. Editrice Compositori, pp. 461–468.
Luisi, P. L. (2001). Are micelles and vesicles chemical equilibrium systems? J. Chem. Educ., 78, 380–384.Google Scholar
Luisi, P. L. (2003a). Contingency and determinism. Phil. Trans. R. Soc. Lond., A, 361, 1141–1147.Google Scholar
Luisi, P. L. (2006). The Emergence of Life. From Chemical Origins to Synthetic Biology. Cambridge University Press.
Luisi, P. L. (2007). Chemical aspects of synthetic biology. Chemistry and Biodiversity, 4, 603–621.Google Scholar
Luisi, P. L. and Chiarabelli, C. (2011). Chemical Synthetic Biology. Hoboken, NJ: John Wiley & Sons.
Luisi, P. L. and Magid, L. (1986). Solubilization of enzymes and nucleic acids in hydrocarbon micelar solutions. Crit. Rev. Biochem., 20, 409–474.Google Scholar
Luisi, P. L. and Stano, P., eds. (2011). The Minimal Cell. The Biophysics of Cell Compartment and the Origin of Cell Functionality. Dordrecht: Springer.
Luisi, P. L., and Straub, B., eds. (1984). Reverse Micelles. Plenum Press.
Luisi, P. L. and Varela, F. J. (1990). Self-replicating micelles – a chemical version of minimal autopoietic systems. Orig. Life Evol. Biosph., 19, 633–643.Google Scholar
Luisi, P. L. and Walde, P., eds. (2000). Giant Vesicles, Perspectives in Supramolecular Chemistry. New York: John Wiley & Sons.
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, 1384–1389.Google 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, 209–246.Google Scholar
Luisi, P. L., Scartazzini, R., Haering, G., and Schurtenberger, P. (1990). Organogels from water-in-oil microemulsions. Colloid Polymer Sci., 268, 356–374.Google Scholar
Luisi, P. L., Lazcano, A., and Varela, F. (1996). In Rizzotti, M., ed., Defining Life: the Central Problem in Theoretical Biology. University of Padova, pp. 149–165.
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): 1759–1777.Google Scholar
Luisi, P. L., Stano, P., Rasi, S., and Mavelli, F. (2004). A possibile route to prebiotic vesicle reproduction. Artificial Life, 10, 297–308.Google Scholar
Luisi, P. L., Ferri, F., and Stano, P. (2006). Approaches to semi-synthetic minimal cells: a review. Naturwissenschaften, 93, 1–13.Google Scholar
Luisi, P. L., Allegretti, M., de Souza, T. P., Steiniger, F., Fahr, A., and Stano, P. (2010). Spontaneous protein overcrowding in liposomes: a new vista for the origin of cellular metabolism. Chem. Biochem., 11, 1989–1992.Google Scholar
Luisi, P. L., Stano, P., and Chiarabelli, C., eds. (2014). Synthetic Biology. Current Opinion in Chemical Biology, 22, pp. v–vii (Editorial overview) and 1–162.Google Scholar
Luther, A., Brandsch, R., and von Kiedrowski, G. (1998). Surface promoted replication and exponential amplification. Nature, 396, 245–248.Google 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, 540–551.Google Scholar
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, 5058–5063.Google Scholar
Machy, P. and Leserman, L. (1987). Liposomes in Cell Biology and Pharmacology. London: John Libbey and Co., Ltd.
Madeira, V. M. C. (1977). Biochim. Biophys. Acta, 499, 202–211.
Maden, B. and Monro, R. E. (1968). Ribosome-catalyzed peptidyl transfer – effects of cations and pH value. European Journal of Biochemistry, 6, 309.Google Scholar
Mader, S. S. (1996). Biology, edn. W. C. Brown Publisher.
Maestro, M. and Luisi, P. L. (1990). A simplified thermodynamic model for protein uptake by reverse micelles. In Mittal, K. L., ed., Surfactants in Solution, Vol. 9. Plenum.
Malyshev, D. A., Dhami, K., Quach, H. T., Lavergne, T., and Ordoukhanian, P. (2012). Efficient and sequence-independent replication of DNA containing a third base pair establishes a functional six-letter genetic alphabet. Proc. Natl. Acad. Sci. USA, 109 (30): 12005–12010.Google Scholar
Mandelbrot, B. (1982). Fractal Geometry of Nature. Fenn and Company.
Mandell, D. J. et al. (2015). Biocontainment of genetically modified organisms by synthetic protein design. Nature, 518, 55–60. http://dx.doi.org/10.1038/nature14121.Google Scholar
Mangiarotti, G. and Chiaberge, S. (1997). Reconstitution of functional eukaryotic ribosomes from Dictyostelium discoideum ribosomal proteins and RNA. The Journal of Biological Chemistry, 272, 19682–19687.Google Scholar
Mansy, S. S. and Szostak, J. W. (2009). Reconstructing the emergence of cellular life through the synthesis of model protocells. Cold Spring Harbor Symp. Quant. Biol., 74, 47–54.Google Scholar
Mansy, S. S., Schrum, J. P., Krishnamurthy, M., Tobé, S., Treco, D. A., and Szostak, J. W. (2008). Template-directed synthesis of a genetic polymer in a model protocell. Nature. doi:10.1038/nature07018.
Margulis, L. (1993). Symbiosis in Cell Evolution. Freeman.
Margulis, L. and Sagan, D. (1995). What is Life? Weidenfeld and Nicholson.
Mariani, P., Luzzati, V., and Delacroix, H. (1988). Cubic phases of lipid-containing systems: structure analysis and biological implications. J. Mol. Biol., 204, 165–189.Google Scholar
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, 6746–6758.Google 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, 626–629.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, 277–295.Google Scholar
Martinek, K., Levashov, A. V., Klyachko, N., Khmelnttski, Yu. L., and Berezin, I. V. (1986). Micellar enzymology. Eur. J. Biochem., 155, 453–468.Google Scholar
Martini, L. and Mansy, S. S. (2011). Cell-like systems with riboswitch controlled gene expression. Chem. Commun., 47, 10734–10736.Google Scholar
Mascolo, R. (2011). L'emergere della biologia della cognizione. La complessità della vita di Humberto Maturana Romecín. Aracne Editrice, Rome.
Mason, S. F. and Tranter, G. E. (1983). The parity violating energy difference between enantiomeric molecules. Chem. Phys. Lett., 94, 34.Google Scholar
Masters, J. R. (2002). HeLa cells 50 years on: the good, the bad and the ugly. Nat. Rev. Cancer, 2, 315–319.Google Scholar
Matsubayashi, H., Kuruma, Y., and Ueda, T. (December 2014). Cell-free synthesis of SecYEG translocon as the fundamental protein transport machinery. Orig. Life Evol. Biosph., 44(4): 331–334.Google 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, 4829–4837.Google Scholar
Matthews, C. N. (1975). The origin of proteins, heteropolypeptides from hydrogen cyanide and water. Origin of Life, 6, 155–163.Google Scholar
Mattman, L. H. (1992). Cell Wall Deficient Forms: Stealth Phatogens. Boca Raton, FL: CRC Press.
Maturana, H. and Varela, F. (1980). Autopoiesis and Cognition: The Realization of the Living. Reidel.
Maturana, H. R. and Varela, F. J. (1998). The Tree of Knowledge (Based on revised edn. of 1992). Shambala. (First edn. 1984, El árbol del conocimiento. Bases biológicas del entendimiento humano. First English edn. 1987).
Maturana, H., Lettvin, J., McCulloch, W., and Pitts, W. (1960). Life and cognition. Gen. Physiol., 43, 129–175.Google Scholar
Mavelli, F. (2004). Theoretical investigations on autopoietic replication mechanisms. ETH-Z Dissertation Nr. 15218, Zurich.
Mavelli, F. (2012). Stochastic simulations of minimal cells: the Ribocell model. BMC bioinformatics, 13(S 4): S 10. doi:10.1186/1471–2105–13-S4-S10.Google Scholar
Mavelli, F. and Luisi, P. L. (1996). Autopoietic self-reproducing vesicles: a simplified kinetic model. J. Phys. Chem., 100, 16600–16607.Google Scholar
Mavelli, F., Altamura, E., Cassidei, L., and Stano, P., (2014). Recent theoretical approaches to minimal artificial cells. Entropy, 16, 2488–2511.Google Scholar
Maynard-Smith, J. and Szathmáry, E. (1995). The Major Transitions in Evolution. Oxford University Press.
Maynard-Smith, J. and Szathmáry, E. (1999). The Origins of Life. Oxford University Press.
Mayr, E. (1974a). Teleological and teleonomic: a new analysis. Boston Studies in the Philosophy of Science, XIV, 91–117.Google Scholar
Mayr, E. (1974b). The multiple meanings of teleological. In Mayr, Ernst, Toward a New Philosophy of Biology: Observations of an Evolutionist. Cambridge, MA: Harvard University Press, 1988, pp. 44–45.
Mayr, E. (1988). The limits of reductionism. Nature, 331, 475.Google Scholar
Mayr, E. (1992). The idea of teleology. Journal of the History of Ideas, 53(1): 117–135.Google 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, 153–166.Google Scholar
McLaughlin, B. P. (1992). The rise and fall of British emergentism. In Beckermann, A., Flohr, H. and Kim, J., eds., Emergence or Reduction: Essays on the Prospects of Nonreductive Materialism, Library edn. de Gruyter, pp. 49–53.
Meier, C. A. (1992). Wolfgang Pauli und C. G. Jung, Ein Briefwechsel. Springer Verlag.
Menger, F. (1991). Groups of organic molecules that operate collectively. Angew. Chem. Int. Ed. Engl., 30, 1086–1099.Google Scholar
Merleau-Ponty, M. (1967). The Structure of Behaviour. Beacon.
Michalodimitrakis, K. and Isalan, M. (2009). Engineering prokaryotic gene circuits. FEMS Microbiol., 33(1): 27–37.Google Scholar
Mill, J. S. (1872). System of Logic, edn. Longmans, Green, Reader and Dyer.
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, 877–886.Google Scholar
Miller, D. M. and Gulbis, J. M. (2015). Engineering protocells: prospects for self-assembly and nanoscale production lines. Life, 5(2): 1019–1053.Google Scholar
Miller, M. B. and Basler, B. L. (2001). Quorum sensing in bacteria. Ann. Rev. Microbiol., 55, 165–199.Google Scholar
Miller, S. L. (1953). Production of amino acids under possible primitive Earth conditions. Science, 117, 2351–2361.Google Scholar
Miller, S. L. (1998). The endogenous synthesis of organic compounds. In Brack, A., ed., The Molecular Origin of Life. Cambridge University Press.
Miller, S. L. and Bada, J. (1988). Submarine hot springs and the origin of life. Nature, 334, 609–611.Google Scholar
Miller, S. L. and Bada, J. (1991). Extraterrestrial synthesis. Nature, 350, 388–389.Google Scholar
Miller, S. L. and Cleaves, H. J. (2007). Prebiotic chemistry on the primitive Earth. In Rigoutsos, I. and Stephanopoulos, G., eds., Systems Biology. Volume I, Genomics. Oxford – New York: Oxford University Press.
Miller, S. L. and Lazcano, A. (1995). The origin and early evolution of life: prebiotic chemistry, the pre-RNA world, and time. J. Mol. Evol., 41, 689–692.Google Scholar
Miller, S. L. and Parris, M. (1964). Nature, 204, 1248–1250.
Mingers, J. (1992). The problems of social autopoiesis. Int. J. Gen. Syst., 21, 229–236.Google Scholar
Mingers, J. (1995). Self-Producing Systems: Implications and Applications of Autopoiesis. Plenum Press.
Mingers, J. (1997). A critical evaluation of Maturana's constructivist family therapy. Syst. Practice, 10(2): 137–151.Google Scholar
Miranda, M., Amicarelli, F., Poma, A., Ragnelli, A. M., and Arcadi, A. (1988). Biochim. Biophys. Acta, 966, 276–286.
Mojzsis, S. J., Harrison, T. M., and Pidgeon, R. T. (2001). Oxygen-isotope evidence from ancient zircons for liquid water at the Earth's surface 4,300 Myr ago. Nature, 409(6817):178–181.Google Scholar
Monod, J. (1971). Chance and Necessity. A. A. Knopf.
Monro, R. and Marcker, K. A. (1967). Ribosome-catalysed reaction of puromycin with a formylmethionine-containing oligonucleotide. Journal of Molecular Biology, 25, 347–350.Google Scholar
Morgan, C. L. (1923). Emergent Evolution. William and Norgate.
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.Google Scholar
Morowitz, H. J. (1967). Biological self-replicating systems. Prog. Theor. Biol., 1, 35–58.Google Scholar
Morowitz, H. J. (1974). Manufacturing a living organism. Hospital Practice, 9, 210–215.Google Scholar
Morowitz, H. J. (1992). Beginnings of Cellular Life. Yale University Press.
Morowitz, H. J., Deamer, D. W., and Smith, T. (1991). Biogenesis as an evolutionary process. J. Mol. Evol., 33, 207–208.Google Scholar
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, 395–399.Google Scholar
Morowitz, H. J., Kostelnik, J. D., Yang, J., and Cody, G. D. (2000). The origin of intermediary metabolism. Proc. Natl. Acad. Sci. USA, 97, 7704–7709.Google Scholar
Mossa, G., Di Giulio, A., Dini, L., and Finazzi-Agrò, A. (1989). Biochim. Biophys. Acta, 986, 310–314.
Müller, D., Pitsch, S., Kittaka, A., Wagner, E., Wintner, C. E., and Eschenmoser, A. (1990). Chemie von α-Aminonitrilen. Aldomerisierung von Glycolaldehyd-phosphat zu racemischen Hexose-2,4,6-triphosphaten und (in Gegenwart von Formaldehyd) racemischen Pentose-2,4-diphosphaten: rac-Allose-2,4,6-triphosphat und rac-Ribose-2,4-diphosphat sind die Reaktionshauptprodukte. Helvetica Chimica acta, 73(5): 1410–1468.Google Scholar
Murtas, G., Kuruma, Y., Bianchini, P., Diaspro, A., and Luisi, P. L. (2007). Protein synthesis in liposomes with a minimal set of enzymes. Biochem Biophys Res Comm., 363: 12–17.Google Scholar
Mushegian, A. (1999). The minimal genome concept. Curr. Opin. Genetics Develop., 9, 709–714.Google Scholar
Mushegian, A. (2005). Protein content of minimal and ancestral ribosome. RNA Society, 11, 1400–1406.Google Scholar
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, 10268–10273.Google Scholar
Nagel, E. (1961). The Structure of Science. Harcourt.
Nakajima, T., Yabushita, Y., and Tabushi, I. (1975). Amino acid synthesis through biogenic CO2 fixation. Nature, 256, 60–61.Google Scholar
Nakashima, T., Toyota, H., Urabe, I., and Yomo, T. (2007). Effective selection system for experimental evolution of random polypeptides towards DNA-binding protein. J. Bioscence and Bioengineering, 103, 155–160.Google Scholar
Naoi, M., Naoi, M., Shimizu, T., Malviya, A. N., and Yagi, K. (1977). Permeability of amino acids into liposomes. Biochim. Biophys. Acta, 471, 305–310.Google Scholar
Nelson, K. E., Levy, M., and Miller, S. L. (2000). Peptide nucleic acids rather than RNA may have been the first genetic molecule. Proc. Natl. Acad. Sci. USA, 97(8): 3868–3871.Google Scholar
Neumann, J. von, and Burks, A., eds. (1966). Theory of Self-Reproduction Automata. University of Illinois Press.
Newport, J. (1987). Nuclear reconstitution in vitro: stages of assembly around protein–free DNA. Cell, 48, 205–217.Google Scholar
Nicolis, G. and Prigogine, I. (1977). Self-Organization in Nonequilibrium Systems. From Dissipative Structures to Order Through Fluctuations. New York: John Wiley & Sons.
Nierhaus, K. H. and Montejo, V. (1973). Protein involved in peptidyltransferase activity of Escherichia-coli ribosomes. Proc. Natl. Acad. Sci. USA, 70, 1931–1935.Google Scholar
Nierhaus, K. H. and Dohme, F. (1974). Total reconstitution of functionally active 50S ribosomal subunits from Escherichia coli. Proc. Natl. Acad. Sci. USA, 71, 4713–4717.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, 920–930.Google Scholar
Noble, D. (2006). The Music of Life. Oxford University Press.
Noireaux, V. and Libchaber, A. (2004). A vesicle bioreactor as a step toward an artificial cell assembly. Proc. Natl. Acad. Sci. USA, 101, 17669–17674.Google Scholar
Noireaux, V., Bar-Ziv, R., and Libchaber, A. (2003). Principles of cell-free genetic circuit assembly. Proc. Natl. Acad. Sci. USA, 100, 12672–12677.Google Scholar
Nomura, S. M., Yoshikawa, Y., Yoshikawa, K., et al. (2001). Towards proto-cells: “primitive” lipid vesicles encapsulating giant DNA and its histone complex. Chem. Bio. Chem., 6, 457–459.Google Scholar
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, 245–246.Google Scholar
Nomura, S. M., Tsumoto, K., Hamada, T., et al. (2003). Gene expression within cell-sized lipid vesicles. Chem. Bio. Chem., 4, 1172–1175.Google Scholar
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, 915–924.Google Scholar
Noyes, R. M. (1989). Some models of chemical oscillators. J. Chem. Educ., 66, 190–191.Google Scholar
Nucara, L. (2014). La Filosofia di Humberto Maturana. Firenze: Casa Editrice Le Lettere.
Oberholzer, T. and Luisi, P. L. (2002). The use of lipsomes for constructing cell models. J. Biol. Phys., 28, 733–744.Google Scholar
Oberholzer, T., Albrizio, M., and Luisi, P. L. (1995a). Polymerase chain reaction in liposomes. Curr. Biol., 2, 677–682.Google 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, 250–257.Google Scholar
Oberholzer, T., Nierhaus, K. H., and Luisi, P. L. (1999). Protein expression in liposomes. Biochem. Biophys. Res. Commun., 261, 238–241.Google Scholar
O'Connor, T. (1994). Emergent properties. Am. Phil. Q., 31, 91–104.Google Scholar
Okada, M. and Ohno, H. (1972). Assembly mechanism of tobacco mosaic virus particle from its ribonucleic acid and protein. Molec. Gen. Genetics, 114, 205–213.Google Scholar
Okasha, S. (2006). Evolution and the Levels of Selection. Oxford University Press.
Olsson, U. and Wennerstrom, H. (2002). On the ripening of vesicle dispersions. J. Phys. Chem. B, 106, 5135–5138.Google Scholar
O'Malley, M. A., Powell, A., Davies, J. F., and Calvert, J. (2007). BioEssays, 30, 57–65.
Oparin, A. I. (1924). Proiskhozhdenie Zhisni. Moskowski Rabocii.
Oparin, A. I. (1938). Origin of Life. McMillan.
Oparin, A. I. (1953). The Origin of Life. Dover Publications.
Oparin, A. I. (1957). The Origin of Life on Earth, edn. Academic Press.
Oparin, A. I. (1961). Life: Its Nature, Origin and Development. Oliver and Boyd.
Oppenheim, P. and Putnam, H. (1958). The unity of science as a working hypothesis. In Feigl, H., Maxwell, G. and Scriven, M., eds., Minnesota Studies in the Philosphy of Science. University of Minnesota Press, pp. 3–36.
Orgel, L. E. (1968). Evolution of the genetic apparatus. J. Mol. Biol., 38, 381–393.Google Scholar
Orgel, L. E. (1973). The Origins of Life. New York: John Wiley & Sons.
Orgel, L. E. (1994). The origin of life on the Earth. Sci. Amer., 271(4): 53–61.Google Scholar
Orgel, L. E. (2000a). A simpler nucleic acid. Science, 290(5495): 1306–07.Google Scholar
Orgel, L. E. (2000b). Self-organizing biochemical cycles. Proc. Natl. Acad. Sci. USA, 97, 12503–12507.Google Scholar
Orgel, L. E. (2003). Some consequences of the RNA world hypothesis. Orig. Life Evol. Biosph., 33, 211–218.Google Scholar
Orgel, L. E. (2004). Prebiotic chemistry and the origin of the RNA world. Crit Rev Biochem Mol Biol., 39, 99–123.Google Scholar
Oró, J. (1960). Synthesis of adenine from ammonium cyanide. Biochem. Bioph. Res. Commun., 2, 407–412.Google Scholar
Oró, J. (1961). Amino acid synthesis from hydrogen cyanide under possible primitive Earth conditions. Nature, 190, 442–443.Google Scholar
Oró, J. (1994). Early chemical stages in the origin of life. In Bengtson, S., ed., Early Life on Earth: Nobel Symposium n. 84. New York: Columbia University Press, pp. 48–59.
Oró, J. (2002). Historical understanding of life's origin. In Schopf, J. W., ed., Life's Origin, the Beginnings of Biological Evolution. University of California Press, pp. 7–41.
Oró, J. and Kimball, A. P. (1961). Synthesis of purines under possible primitive Earth conditions. 1. Adenine from hydrogen cyanide. Arch. Biochem. Biophys., 94, 221–227.Google 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.Google 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.Google Scholar
Ourisson, G. and Nakatani, Y. (1999). Origin of cellular life: molecular foundations and new approaches. Tetrahedron, 55, 3183–3190.Google Scholar
Ousfouri, S., Stano, P., and Luisi, P. L. (2005). Condensed DNA in lipid microcompartments. J. Phys. Chem. B., 109, 19929–19935.Google Scholar
Pääbo, S. (1993). Ancient DNA. Scientific American, 269(5): 60–66.Google Scholar
Palazzo, G. and Luisi, P. L. (1992). Solubilization of ribosomes in reverse micelles. Biochem. Biophys. Res. Commun., 186, 1546–1552.Google Scholar
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, edn (1986). Lincoln-Rembrandt Publishing.
Palyi, G., Zucchi, C., and Caglioti, L., eds. (2002). Fundamentals of Life. Elsevier.
Pantazatos, D. P. and McDonald, R. C. (1999). Directly observed membrane fusion between oppositely charged phospholipid bilayers. Membrane Biol., 170, 27–38.Google Scholar
Papahadjopoulos, D., Lopez, N., and Gabizon, A. (1989). Drug delivery by liposomes. In Lopez-Berenstein, G. and Fidler, I. J., eds., Liposomes in Therapy of Infectious Diseases and Cancer. Alan Riss Inc., pp. 135–154.
Parens, E., Johnston, J., and Moses, J. (2008). Ethics. Do we need “synthetic bioethics”? Science, 321 (5895): 1449.Google Scholar
Park, J. H. and Lee, S. Y. (2008). Towards systems metabolic engineering of microorganisms for amino acid production. Curr. Opin. Biotechnol., 19(5): 454–460.Google Scholar
Parker, E. T., Zhou, M., Burton, A. S., Glavin, D. P., Dworkin, J. P., et al. (2014). A plausible simultaneous synthesis of amino acids and simple peptides on the primordial Earth. Proc. Natl. Acad. Sci. USA, 108(12): 5526–4431.Google Scholar
Parsons, P. (1996). Dusting off panspermia. Nature, 383, 221–222.Google Scholar
Pascal, R., Boiteax, L., and Commeyras, A. (2005). From the prebiotic synthesis of amino acids towards a primitive translation apparatus. Top Curr. Chem., 259–269.
Patel, B. H., Percivalle, C., Ritson, D. J., Duffy, C. D., and Sutherland, J. D. (2015). Nature Chem., 7, 301–307.
Paul, N. and Joyce, G. F. (2002). A self-replicating ligase ribozyme. Proc. Natl. Acad. Sci. USA, 99, 12733–12740.Google Scholar
Paul, N. and Joyce, G. F. (2004). Minimal self-replicating systems. Curr. Opin. Chem. Biol., 8, 634–639.Google Scholar
Paul, N., Springsteen, G., and Joyce, G. F. (2006). Conversion of a ribozyme to a deoxyribozyme through in vitro evolution. Chem. & Biol., 13, 329–338.Google Scholar
Pereto, J., Lopez-Garcia, P., and Moreira, D. (2004). Ancestral lipid biosynthesis and early membrane evolution. Trends Biochem. Sci., 29, 469–477.Google Scholar
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, 1244–1251.Google Scholar
Pfammatter, N., Hochköppler, A., and Luisi, P. L. (1992). Solubilization and growth of Candida pseudotropicalis in water-in-oil microemulsions. Biotechnol. Bioeng., 40, 167–172.Google Scholar
Pfüller, U. (1986). Mizellen, Vesikeln, Mikroemulsionen. Springer Verlag.
Piaget, J. (1967). Biologie et connaissance. Gallimard.
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.Google Scholar
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.Google Scholar
Pileni, M. P. (1981). Photoelectron transfer in reverse micelles – photo-reduction of cytochrome-c. Chem. Phys. Lett., 81, 603–605.Google Scholar
Pizzarello, S. and Cronin, J. R. (2000). Non-racemic amino acids in the Murray and Murchison meteorites. Geochim. Cosmochim. Acta, 64, 329–338.Google Scholar
Pizzarello, S. and Weber, A. L. (2004). Prebiotic amino acids as asymmetric catalysts. Science, 303, 1151.Google Scholar
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, 225–236.Google Scholar
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, 239–248.Google Scholar
Poerksen, B. (2004). The Certainty of Uncertainty, Dialogues Introducing Constructivism. Imprint Academic.
Pohorille, A. and Deamer, D. (2002). Artificial cells: prospects for biotechnology. Trends Biotech., 20, 123–128.Google Scholar
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.Google Scholar
Pollack, A. (2014). Scientists add letters to DNA's alphabet, raising hope and fear. New York Times, May 7.
Ponce de Leon, S. and Lazcano, A. (2003). Panspermia – true or false? Lancet, 362, 406–407.Google Scholar
Ponnamperuma, C. and Peterson, E. (1965). Peptide synthesis from aminoacids in aqueous solution. Science, 147, 1572.Google Scholar
Popa, R. (2004). Between Necessity and Probability: Searching for the Definition and Origin of Life. Springer Verlag.
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.Google 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, 8437–8440.Google Scholar
Powner, M. W. and Sutherland, J. D. (2010). Phosphate-mediated interconversion of ribo- and arabino-configured prebiotic nucleotide intermediates. Angew Chem. Int. Ed., 49, 4641–4643.Google Scholar
Powner, M. W., Gerland, B., and Sutherland, J. D. (2009). Synthesis of activated pyrimidine ribonucleotides in prebiotically plausible conditions. Nature, 459, 239–242.Google Scholar
Powner, M. W., Sutherland, J. D., and Szostak, J. W. (2010). Chemoselective multicomponent one-pot assembly of purine precursors in water. J. Am. Chem. Soc., 24, 16677–16688.Google Scholar
Pozzi, G., Birault, V., and Werner, B. (1996). Single-chain polyprenyl phosphates form primitive membranes. Angew. Chem. Int. Ed. Engl., 35, 177–179.Google Scholar
Prigogine, I. (1997). The End of Certainty-Time, Chaos and the New Laws of Nature. Free Press.
Prigogine, I. and Lefever, R. (1968). Symmetry breaking instabilities in dissipative systems. J. Chem. Phys., 48, 1695–1700.Google Scholar
Prijambada, I.D., et al. (1996). Solubility of artificial proteins with random sequences. Febs Letters. 382, 21–25.Google Scholar
Primas, H. (1985). Can chemistry be reduced in physics? Chem. Uns. Zeit, 19, 160.Google Scholar
Primas, H. (1993). In Fischer, E. P., ed., Neue Horizonte 92/93: Ein Forum der Naturwissenschaften. München: Piper.
Primas, H. (1998). Emergence in exact natural sciences. Acta Politechnica Scand., 91, 86–87.Google Scholar
Pross, A. (2005). On the chemical nature and origin of teleonomy. Origin of Life & Evol. Biosphere, 35, 384–394.Google Scholar
Pryer, W. (1880). Die Hypothesen über den Ursprung des Lebens. Berlin.
Purrello, R. (2003). Lasting chiral memory. Nature Mater., 2, 216–217.Google Scholar
Pyun, J., Zhou, X.-Z., Drockenmuller, E., and Hawker, C. J. (2003). Mater. Chem., 13, 2653.
Qian, L. and Winfree, E. (2011). Scaling up digital circuit computation with DNA strand displacement cascades, Science, 332, 1196–1201. doi:10.1126/science.1200520.Google Scholar
Quack, M. (2002). Angew. Chem., 41, 4618–4630.
Quack, M. and Stohner, J. (2003a). Combined multidimensional anharmonic and parity violating effects in CDBrClF. J. Chem. Phys., 119, 11228–40.Google 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, 375–376.Google Scholar
Quack, M. and Stohner, J. (2014). The concept of law models in chemistry. European review, 22, S 50–86.Google Scholar
Raab, W. (1988). Ärtzliche Kosmetologie, 18, 213–224.
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, 6105–6109.Google Scholar
Rajamani, S., Vlassov, A., Benner, S., Coombs, A., Olasagasti, F., and Deamer, D. (2007). Lipid-assisted synthesis of RNA-like polymers from mononucleotides. Orig. Life Evol. Biosph. doi:10.1007/s11084-007-9113-2.
Ramundo-Orlando, A., Arcovito, C., Palombo, A., Serafino, A. L., and Mossa, G. (1993). J. Liposome Res., 3, 717–724.
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.Google Scholar
Rao, M., Eichberg, J., and Oró, J. (1987). Synthesis of phosphatidylethanolamine under possible primitive earth conditions. J. Mol. Evol., 25, 1–6.Google Scholar
Rasi, S., Mavelli, F., and Luisi, P. L. (2003). Cooperative micelle binding and matrix effect in oleate vesicle formation. J. Phys. Chem. B, 107, 14068–14076.Google Scholar
Rasi, S., Mavelli, F., and Luisi, P. L. (2004). Matrix effect in oleat-micelles-vesicles transformations. Orig. Life Evol. Bioph., 34, 215–24.Google Scholar
Rathman, J. F. (1996). Micellar catalysis. Curr. Opin. Coll. Interf. Sci., 1, 514–518.Google Scholar
Rebek, J. (1994). A template for life. Chem. Br., 30, 286–290.Google Scholar
Recordati, G. and Bellini, T. G. (2004). A definition of internal constancy and homeostasis in the context of non-equilibrium thermodynamics. Experimental Physiology, 89(1): 27–38.Google Scholar
Reichenbach, H. (1978). The aims and methods of physical knowledge. In Reichenbach, M. and Cohen, R. S., eds., Hans Reichenbach: Selected Writings 1909–53. Translated by Schneewind, E. H.. Reidel, pp. 81–225.
Reszka, R. (1998). Liposomes as drug carrier for diagnostics, cytostatics and genetic material. In Diederichs, J. E. and Müller, R. H., eds., Future Strategies for Drug Delivery with Particulate Systems. Medpharm GmbH Scientific Publishers.
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, 2063–2066.CrossRefGoogle Scholar
Riddle, D. S., Santiago, J. V., Bray-Hall, S. T., Doshi, N., Grantcharova, V. P., Yi, Q., and Baker, D. (1997). Functional rapidly folding proteins from simplified amino acid sequences. Nature, 4, 805–809.Google Scholar
Ringertz, N. R., Krondahl, U., and Coleman, J. R. (1978). Reconstitution of cells by fusion of cell fragments. Experimental Cell Research, 113, 233–246.Google Scholar
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.Google Scholar
Riste, T. and Sherrington, D., eds. (1996). Physics of Biomaterials: Fluctuations, Selfassembly and Evolution (Nato Science Series, Series E, Applied Sciences). Kluwer.
Rizzotti, M., ed. (1996). Defining Life. University of Padua.
Robertson, M. P. and Joyce, G. F. (2012). The origins of the RNA world. Cold Spring Harb Perspect Biol., 4, 5. doi:10.1101/cshperspect.a003608.Google Scholar
Robertson, R. N. (1983). The Lively Membrane. Cambridge University Press.
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, 273–286.Google Scholar
Rodrigo, G., Landrain, T. E., and Jaramillo, A. (2012). De novo automated design of small RNA circuits for engineering synthetic riboregulation in living cells. Proc. Natl. Acad. Sci. USA., 109, 15271–15276. doi:10.1073/pnas.1203831109.Google Scholar
Rodrigo, G., Landrain, T. E., Majer, E., Daròs, J.-A., and Jaramillo, A. (2013). Full design automation of multi-state RNA devices to program gene expression using energy-based optimization. PLoS Comp. Biol., 9, e1003172. doi:10.1371/journal.pcbi.1003172.Google Scholar
Rogerson, M. L., Robinson, B. H., Bucak, S., and Walde, P. (2006). Kinetic studies of the interaction of fatty acis with phosphatidylcholine vesicles (Liposomes). Colloid and Surfaces B: Biointerfaces, 48, 24–34.Google Scholar
Rohl, C. A., Strauss, C. E., Misura, K. M., and Baker, D. (2004). Protein structure prediction using Rosetta. Methods Enzymol., 383, 66–93.Google Scholar
Rojas, N. R. L., et al. (1997). De novo heme proteins from designed combinatorial libraries. Protein Science, 6, 2512–2524.Google Scholar
Rolle, F. (1863). Ch. Darwin's Lehre von der Entstehung der Arten, in ihrer Anwendung auf die Schöpfunggeschichte. J. C. Hermann.
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, 205–222.Google Scholar
Rotello, V., Hong, J. I., and Rebek, J. (1991). Sigmoidal growth in a self-replicating system. J. Am. Chem. Soc., 113, 9422–9423.Google Scholar
Runion, G. E. (1990). The Golden Section. Dale Seymour Publications.
Rushdi, A. I. and Simoneit, B. R. (2001). Lipid formation by aqueous Fischer-Tropf-type synthesis over a temperature range 100 to 400 °C. Orig. Life Evol. Biosph., 31, 103–118.Google Scholar
Sackmann, E. (1978). Dynamic molecular-organization in vesicles and membranes. Ber. Bunsen-Gesell. Phys. Chem., 82, 891–909.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, 826–830.Google Scholar
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, 665–671.Google Scholar
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, 167–176.Google Scholar
Sagan, C. (1985). Cosmos. Ballantine Publishing.
Sagan, C. (1994). The search for extraterrestrial life. Sci. Amer., 271(4): 71–77.Google Scholar
Saito, H., Yamada, A., Ohmori, R., Kato, Y., Yamanaka, T., et al. (2007). Towards constructing synthetic cells: RNA/RNP evolution and cell-free translational systems in giant liposomes. Micro-NanoMechatronics and Human Science, 2007. MHS ’07 (International Symposium on), 286–291.
Samoylov, A. M., Samoylova, T. I., Pustovyy, O. M., et al. (2005). Novel metal cluster isolated from blood are lethal to cancer cells. Cells Tissues Organs, 179(3): 115–124.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, 72–73.Google 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, 121–28.Google Scholar
Scartazzini, R. and Luisi, P. L. (1988). Organogels from lecithins. J. Phys. Chem., 92, 829–833.Google Scholar
Schaerer, A. A. (2002). Conceptual conditions for conceiving life – a solution for grasping its principle, not mere appearances. In Palyi, G., Zucchi, C., and Caglioti, L., eds., Fundamentals of Life. Elsevier, pp. 589–624.
Schafmeister, C. E., et al. (1997). A designed four-helix bundle protein with native like structure. Nature, 4, 1039–1046.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, 8127–8130.Google Scholar
Schopf, J. W. and Klein, C., eds. (1992). In The Proterozoic Atmosphere. Cambridge University Press.
Schopf, J. W. (1993). Microfossils of the early archean apex chert: new evidence of the antiquity of life. Science, 260, 640–646.Google Scholar
Schopf, J. W. (1998). Chemical evolution and the origin of life. In Brack, A., ed., The Molecular Origin of Life. Cambridge University Press.
Schopf, J. W. (2002). Life's Origin. University of California Press.
Schröder, J. (1998). Emergence: non-deducibility or downward causation? Phil. Q., 48, 434–452.Google Scholar
Schulze, H. and Nierhaus, K. H. (1982). Minimal set of ribosomal components for reconstitution of the peptidyltransferase activity. Embo Journal, 1, 609–613.Google 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, 3695–3701.Google 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, 4173–4176.Google Scholar
Schuster, P. and Swetina, J. (1988). Stationary mutant distributions and evolutionary optimization. Bull. Math. Biol., 50, 636–660.Google Scholar
Schwabe, C. (2001). The Genomic Potential Hypothesis, a Chemist's View on the Origin and Evolution of Life. Landes Bioscience.
Schwabe, C. and Warr, G. W. (1984). A polyphyletic view of evolution. The genetic potential hypothesis. Persp Biol. Med., 27, 465–485.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.Google Scholar
Seddon, J. M., Hogan, J. J., Warrender, N. A., and Pebay-Peyroula, E. (1990). Prog. Coll. Polym. Sci., 81, 189–197.
Sela, M., White, F. H. Jr., and Anfinsen, C. B. (1957). Reductive cleavage of disulfide bridges in ribonuclease. Science, 125, 691–692.Google Scholar
Selsis, F. (2000). Modèle d’évolution physico-chimique des atmosphères de planètes telluriques. Application à l'atmosphère primitive terrestre et aux planètes extrasolaires. Ph.D. thesis, Université Bordeaux 1 (France).
Severin, K., Lee, D. H., Kennan, A. J., and Ghadiri, M. R. (1997). A synthetic peptide ligase. Nature, 16(389): 706–709.Google Scholar
Shapiro, R. (1984). The improbability of prebiotic nucleic acid synthesis. Orig. Life, 14(1–4): 565–570.Google Scholar
Shapiro, R. (1986). Origins: a Skeptic's Guide to the Creation of Life on Earth. Summit Books.
Shapiro, R. (1988). Prebiotic ribose synthesis: a critical analysis. Orig. Life Evol. Biosph., 18, 71–85.Google Scholar
Shapiro, R. (1995). The prebiotic role of adenine: a critical analysis. Orig. Life Evol. Biosph., 25, 83–98.Google Scholar
Shapiro, R. (2000). A replicator was not involved in the origin of life. IUBMB Life, 49, 173–176.Google Scholar
Shen, C., Lazcano, A., and Oró, J. (1990a). The enhancement activities of histidyl-histidine in some prebiotic reactions. J. Mol. Evol., 31, 445–452.Google Scholar
Shen, C., Mills, T., and Oró, J. (1990b). Prebiotic synthesis of histidyl-histidine. J. Mol. Evol., 31, 175–179.Google Scholar
Shen, C., Yang, L., Miller, S. L., and Oró, J. (1990c). Prebiotic synthesis of histidine. J. Mol. Evol., 31, 167–174.Google Scholar
Sheng, J., Li, L., Engelhart, A.E., Gan, J., Wang, J., and Szostak, J.W. (2014). Structural insights into the effects of 2′- 5′ linkages on the RNA duplex. Proc. Natl. Acad. Sci. USA, 111(8): 3050–3055. doi:10.1073/pnas.1317799111.Google Scholar
Shermer, M. (2003). Is the universe fine-tuned for life? Sci. Amer., Jan. 23.
Shimizu, Y., Inoue, A., Tomari, Y., Suzuki, T., Yokogawa, T., Nishikawa, K., and Ueda, T. (2001). Cell-Free Translation Reconstituted with Purified Components. Nat. Biotechnol., 19, 751–755.Google Scholar
Shimizu, Y., Kanamori, T., and Ueda, T. (2005). Protein Synthesis by Pure Translation. Systems Methods, 36, 299–304.Google Scholar
Shimoyama, A. and Ogasawara, R. (2002). Dipeptides and diketopiperazines in the Yamato-791198 and Murchison carbonaceous chondrites. Orig. Life Evol. Biosph., 32(2): 165–179.Google Scholar
Shimkets, L. J. (1998). Structure and sizes of genomes of the archaea and bacteria. In De Bruijn, F. J., Lupskin, J. R., and Weinstock, G. M., eds., Bacterial Genomes: Physical Structure and Analysis. Kluwer, pp. 5–11.
Shiner, E. K., Rumbaugh, K. P., and Williams, S. C. (2005). Interkingdom signaling: deciphering the language of acyl homoserine lactones. FEMS Microbiol. Rev., 29, 935–947.Google Scholar
Shohda, K. and Sugawara, T. (2006). DNA polymerization on the inner surface of a giant liposome for synthesizing an artificial cell model. Soft Matter, 2, 402–408.Google Scholar
Sievers, D. and von Kiedrowski, G. (1994). Self-replication of complementary nucleotide-based oligomers. Nature, 369, 221–224.Google Scholar
Sievers, D., Achilles, T., Burmeister, J., et al. (1994). Molecular replication – from minimal to complex systems. In Fleischacker, G., Colonna, S. and Luisi, P. L., eds., Self-Production of Supramolecular Structures. Kluwer Publishers.
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, 14676–14683.Google Scholar
Silverman, J. A., Balakrishnan, R., and Harbury, P. B. (2001). Reverse engineering the (β/α)8 barrel fold. Proc. Natl. Acad. Sci. USA, 98, 3092–3097.Google Scholar
Simpson, G. G. (1973). Added comments on “The non-prevalence of humanoids.” In Sagan, C., ed., Communication with Extraterrestrial Intelligence. MIT Press, pp. 362–364.
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, 15440–15445.Google Scholar
Smith, R. S. and Iglewski, B. H. (2003). P. aeruginosa quorum sensing systems and virulence. Curr. Opin. Microbiol., 6, 56–60.Google Scholar
Soai, K., ed. (2008). Amplification of Chirality (Topics in Current Chemistry, vol. 284). Springer.
Soga, H., Fuji, S., Yomo, T., Kato, Y., Watanabe, H., and Matsuura, T. (2014). In vitro membrane protein synthesis inside cell-sized vesicles reveals the dependence of membrane protein integration on vesicle volume. ACS Synth. Biol., 3(6): 372–379.Google Scholar
Solomon, B. and Miller, I. R. (1976). Interaction of glucose oxidase with phospholipid vesicles. Biochim. Biophys. Acta, 455, 332–342.Google Scholar
Sommer, A. P. and Wickramasinghe, N. C. (2005). Functions and possible provenance of primordial proteins-part II: Microorganism aggregation in clouds triggered by climate change. J. Proteome Res., 4, 180–184.Google Scholar
Spang, A., Saw, J. H., Jørgensen, S. L., Zaremba-Niedzwiedzka, K., Martijn, J., et al. (2015). Complex archaea that bridge the gap between prokaryotes and eukaryotes. Nature, 521(7551): 173–179. doi: 10.1038/nature14447.Google Scholar
Sperry, R. W. (1945). Journal of Neurophysiology, 8, 15.
Sperry, R. W. (1986). Discussions: macro- versus microdeterminism. Phil. Sci., 53, 265–270.Google Scholar
Spirin, A. (1986). Ribosome Structure and Protein Synthesis. Benjamin Cummings Publishing.
Sprinzak, D. and Elowitz, M. B. (2005). Reconstruction of genetic circuits. Nature, 438(7067): 443–448.Google Scholar
Stano, P. and Luisi, P. L. (2007). Basic questions about the origins of life: proceedings of the Erice International School of Complexity. Origins of Life and Evolution of Biospheres, 37, 303–307.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.Google Scholar
Stano, P., Wehrli, E., and Luisi, P.L. (2006). Insights on the oleate vesicles self-reproduction. J. Physics Condensed Matter, 18 S2231–S2238.Google Scholar
Stano, P., Carrara, P., Kuruma, Y., de Souza, T. P., and Luisi, P. L. (2011). Compartmentalized reactions as a case of soft-matter biotechnology: synthesis of proteins and nucleic acids inside lipid vesicles. J. Mater. Chem., 21, 18887–18902.Google Scholar
Stano, P., Rampioni, G., Damiano, L., D'Angelo, F., Carrara, P., Leoni, L., and Luisi, P. L. (2014). Towards the engineering of chemical communication between semi-synthetic and natural cells. In Cagnoni, S., Mirolli, M. and Villani, M., eds., Evolution, Complexity and Artificial Life. Dordrecht: Springer, pp. 91–104.
Stetter, K. O. (1998). Hyperthermophiles and their possible role as ancestors of modern life. In Brack, A., ed., The Molecular Origin of Life. Cambridge University Press.
Stocks, P. G. and Schwartz, A. W. (1982). Basic nitrogen-heterocyclic compounds in the Murchison meteorite. Geochim. Cosmochim. Acta, 46, 309–315.Google Scholar
Strogatz, S. H. (1994). Non Linear Dynamics and Chaos, With Applications to Physics, Biology, Chemistry, and Engineering. Perseus Book Group.
Strogatz, S. (2001). Exploring complex networks. Nature, 410, 268–276.Google Scholar
Stryer, L. (1975). Biochemistry. Freeman and Co.
Summers, D. P. and Chang, S. (1993). Prebiotic ammonia from iron(II) reduction of nitrite on the early earth. Nature, 365, 630–633.Google Scholar
Summers, D. P. and Lerner, N. R. (1998). Ammonia from iron(II) reduction of nitrite and the Strecker synthesis: do iron(II) and cyanide interfere with each other? Orig. Life Evol. Biosphere, 28, 1–11.Google Scholar
Sunami, T., Sato, K., Matsuura, T., Tsukada, K., Urabe, I., and Yomo, T. (2006). Femtoliter compartment in liposomes for in vitro selection of proteins. Analytical Biochemistry, 357, 128–136.Google Scholar
Sunami, T., Caschera, F., Morita, Y., Toyota, T., Nishimura, K., Matsuura, T., Suzuki, H., Hanczyc, M. M., and Yomo, T. (2010). Detection of association and fusion of giant vesicles using a fluorescence-activated cell sorter. Langmuir, 26, 15098–15103.Google Scholar
Susskind, L. (2005). The Cosmic Landscape: String Theory and the Illusion of the Intelligent Design. Little Brown.
Sutherland, J. D. (2007). Looking beyond the RNA structural neighborhood for potentially primordial genetic system. Angewandte Chemie Int. Ed., 46, 2354–2356.Google Scholar
Sutherland, J. D., Anastasi, C., Buchet, F. F., Crower, M. A., Parkes, A. L., Powner, M. W., and Smith, J. M. (2007). RNA: prebiotic product, or biotic invention. Chemistry & Biodiversity, 4(4): 721–739.Google Scholar
Swairjo, M. A., Seaton, B. A., and Roberts, M. F. (1994). Biochem. Biophys. Acta, 1191, 354–361.Google Scholar
Szathmáry, E. (2002). Units of evolution and units of life. In Palyi, G., Zucchi, L. and Caglioti, L., eds., Fundamentals of Life. Elsevier SAS, pp. 181–195.
Szostak, J. W., Bartel, D. P., and Luisi, P. L. (2001). Synthesizing life. Nature, 409, 387–390.Google Scholar
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, 638–645.Google Scholar
Takahashi, Y. and Mihara, H. (2004). Construction of a chemically and conformationally self-replicating system of amyloid-like fibrils. Bioorg. Med. Chem., 12, 693–699.Google Scholar
Takakura, K., Toyota, T., and Sugawara, T. (2003). A novel system of self-reproducing giant vesicles. J. Am. Chem. Soc., 125, 8134–8140.Google Scholar
Takakura, K., Yamamoto, T., Kurihara, K., Toyota, T., Ohnuma, K., and Sugawara, T. (2014). Spontaneous transformation from micelles to vesicles associated with sequential conversions of comprising amphiphiles within assemblies. Chem. Commun. (Camb)., 50(17): 2190–2192. doi:10.1039/c3cc47786j.Google Scholar
Tanford, C. (1978). The hydrophobic effect and the organization of living matter. Science, 200, 1012–1018.Google Scholar
Tegmark, M. (2003). Parallel universes. Scientific American, May, 41–51.
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, 744–748.Google Scholar
Thomas, C. F. and Luisi, P. L. (2004). Novel properties of DDAB: matrix effect and interaction with oleate. J. Phys. Chem. B, 108, 11285–11290.Google Scholar
Thomas, C. F. and Luisi, P. L. (2005). RNA selectively interacts with vesicles depending on their size. J. Phys. Chem. B., 109, 14544–14550.Google Scholar
Thompson, E. (2007). Mind in Life. The Belknap Press of the Harvard University Press.
Thompson, E. (2014). Waking, Dreaming, Being: Information and Consciousness in Neuroscience. Columbia University Press.
Thompson, E. and Varela, F. J. (2001). Radical embodiment: neural dynamics and consciousness. Trends Cog. Sci., 5, 418–425.Google Scholar
Torre, P., Keating, C. D., and Mansy, S. S. (2014). Multiphase water-in-oil emulsion droplets for cell-free transcription-translation. Langmuir, 30, 5695–5699.Google Scholar
Traub, P. and Nomura, M. (1968). Structure and function of E. coli ribosomes, V. reconstitution of functionally active 30S ribosomal particles from RNA and proteins. Proc. Nat. Acad. Sci. USA., 59, 737–741.Google Scholar
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, 7225–7228.Google Scholar
Turing, A. (1952). The chemical basis of morphogenesis. Phil. Trans. Royal. Soc. London B, 237, 37.Google Scholar
Ulbricht, W. and Hoffmann, H. (1993). Physikalische Chemie der Tenside. In Kosswig, K., and Stache, H., ed., Die Tenside. Carl Hanser Verlag, pp. 1–114.
Ulman, A. (1996). Formation and structure of self-assembled monolayers. Chem. Rev., 96, 1533–54.Google Scholar
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): 385–395.Google Scholar
Uwin, P. J. R., Webb, R. I., and Taylor, A. P. (1998) Novel nano-organisms from Australian sandstones. Am. Mineralogist, 83, 1541–1550.Google Scholar
Valenzuela, C. Y. (2002). Does biotic life exist? In Palyi, G., Zucchi, C., and Cagiliati, L., eds., Fundamentals of Life. Elsevier, pp. 331–334.
Vancanneyt, M., Schut, F., Snauwaert, C., Goris, J., Swings, J., and Gottschal, J. C. (2001). Sphingopyxis alaskensis sp. Nov, a dominant bacterium from a marine oligotrophic environment. Int. J. Syst. Evol. Microbiol., 51, 73–79.Google Scholar
Van der Gulik, P., Massar, S., Gilis, D., Buhrman, H., and Rooman, M. (2009). The first peptides: the evolutionary transition between prebiotic amino acids and early proteins. Journal of Theoretical Biology, 261(4): 531–553.Google Scholar
Varela, F. J. (1979). Principles of Biological Autonomy. North Holland/Elsevier. Translated in French, see infra (Varela, 1989b).
Varela, F. J. (1989a). Reflections on the circulation of concepts between a biology of cognition and systemic family therapy. Family Process, 28, 15–24.Google Scholar
Varela, F. J. (1989b). Autonomie et Connaissance. Seuil.
Varela, F. J. (1999). Ethical Know-How: Action, Wisdom, and Cognition. Stanford University Press.
Varela, F. J. (2000). El Fenómeno de la Vita. Dolmen Ensayo.
Varela, F. J., Maturana, H. R., and Uribe, R. B. (1974). Autopoiesis: the organization of living system, its characterization and a model. Biosystems, 5, 187–196.Google Scholar
Varela, F. J., Thompson, E., and Rosch, E. (1991). The Embodied Mind: Cognitive Science and Human Experience. Cambridge, MA: MIT Press.
Veomett, G., Prescott, D. M., Shay, J., and Porter, K. R. (1974). Reconstruction of mammalian cells from nuclear and cytoplasmic components separated by treatment with cytochalasin B. Proc. Nat. Acad. Sci. USA., 71, 1999–2002.Google Scholar
Vilanova, C. and Porcar, M. (2014). Table 1: Finalist projects in the iGEM competition, 2006–2013. Nature Biotechnology, 32, 420–424. doi:10.1038/nbt.2899.Google Scholar
Villarreal, L. P. (2009). The source of self: genetic parasites and the origin of adaptive immunity. Ann. N. Y. Acad. Sci., 1178, 194–232.Google Scholar
Villarreal, L. P. (2011). Viral ancestors of antiviral systems. Viruses, 3(10): 1933–1958.Google Scholar
Villarreal, L. P., and Witzany, G. (2010). Viruses are essential agents within the roots and stem of the tree of life. J. Theor. Biol., 262(4): 698–710.Google Scholar
Wächtershäuser, G. (1988). Before enzymes and templates: theory of surface metabolism. Microbiol. Rev., 52, 452–484.Google Scholar
Wächtershäuser, G. (1990a). Evolution of the first metabolic cycles. Proc. Natl. Acad. Sci. USA, 87, 200–204.Google Scholar
Wächtershäuser, G. (1990b). The case for the chemoautotrophic origin of life in the iron–sulfur world. Origin Life Evol. Biosph., 20, 173–176.Google Scholar
Wächtershäuser, G. (1992). Groundworks for an evolutionary biochemistry: the iron–sulfur world. Prog. Biophys. Mol. Biol., 58, 85–201.Google Scholar
Wächtershäuser, G. (1997). The origin of life and its methodological challenge. J. Theor. Biol., 187, 483–494.Google Scholar
Wächtershäuser, G. (2000). Life as we don't know it. Science, 289, 1307–1308.Google Scholar
Waks, M. (1986). Proteins and peptides in water-restricted environments. Proteins, 1, 4–15.Google Scholar
Waks, Z. and Silver, P. A. (2009). Engineering a synthetic dual-organism system for hydrogen production. Appl. Environ. Microbiol., 75(7): 1867–1875.Google Scholar
Walde, P. (2000). Enzymatic reactions in giant vesicles. In Luisi, P. L. and Walde, P., eds., Giant Vesicles, Perspectives in Supramolecular Chemistry. John Wiley & Sons, pp. 297–311.
Walde, P. and Ishikawa, S. (2001). Enzymes inside lipid vesicles: preparation, reactivity and applications. Biomol. Eng., 18, 143–177.Google Scholar
Walde, P. and Mazzetta, B. (1998). Bilayer permeability-based substrate selectivity of an enzyme in liposomes. Biotechnol. Bioeng., 57, 216–219.Google 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(17): 7541–7547.Google 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(26): 11649–11654.Google Scholar
Walde, P., Cosentino, K., Hengel, H., and Stano, P. (2010). Giant vesicles: preparations and applications. ChemBioChem, 11, 848–865.Google Scholar
Walter, K. U., Vamvaca, K., and Hilvert, D. (2005). An active enzyme constructed from a 9-amino acid alphabet. The Journal of Biological Chemistry, 280, 37742–37746.Google Scholar
Wang, J. and Wang, W. (1999). A computational approach to simplifying the protein folding alphabet. Nature Structural Biology, 6, 1033–1038.Google Scholar
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
Weber, W., Lienhart, C., El-Baba, M. D., and Fussenegger, M. (2009). A biotin-triggered genetic switch in mammalian cells and mice. Metabolic Engineering, 11(2): 117–124.Google Scholar
Wei, Y. and Hecht, M. H. (2004). Enzyme-like proteins from an unselected library of designed amino acid sequences. Protein Engineering, Design & Selection, 17, 67–75.Google Scholar
Wei, Y., Liu, T., Sazinskiy, S. L., Moffet, D. A., Pelczer, I., and Hecht, M. H. (2003). Stably folded de novo proteins from a designed combinatorial library. Protein Science, 12, 92–102.Google Scholar
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.
Whitesides, G. M. and Boncheva, M. (2002). Beyond molecules: self-assembly of mesoscopic and macroscopic components. Proc. Natl. Acad. Sci. USA, 99, 4769–4774.Google Scholar
Whitesides, G. M. and Grzybowski, B. (2002). Self-assembly at all scales, Science, 295, 2418–2421.Google Scholar
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, 1312–1319.Google Scholar
Whitfield, J. (2006). In the Beat of a Heart: Life, Energy, and the Unity of Nature. National Academies Press.
Wick, R., Walde, P., and Luisi, P. L. (1995). Autocatalytic self-reproduction of giant vesicles. J. Am. Chem. Soc., 117, 1435–1436.Google Scholar
Wick, R., Angelova, M., Walde, P., and Luisi, P. L. (1996). Microinjection into giant vesicles and light microscopy investigations of enzyme mediated vesicle transformations. Chemistry and Biology, 3, 105–111.Google Scholar
Wieczorek, R., Dorr, M., Chotera, A., Luisi, P. L., and Monnard, P.-A. (2013). Formation of RNA phosphodiester bond by histidine containing dipeptides. ChemBioChem, 14(2): 217–223.Google Scholar
Williams, T. A., Foster, P. G., Cox, C. J., and Embley, T. M. (2013). An archaeal origin of eukaryotes supports only two primary domains of life. Nature, 504, 231–236.Google 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.Google Scholar
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, 6011–6021.Google Scholar
Wilson, T. L. (2001). The search for extraterrestrial intelligence. Nature, 409, 1110–1114.Google Scholar
Wimsatt, W. C. (1972). Complexity and organization. In Schaffner, K. F. and Cohen, R. S., eds., Boston Studies in the Philosophy of Science, Proceedings of the Philosphy of Science Association. Reidel, pp. 67–86.
Wimsatt, W. C. (1976a). Reductionism, levels of organization, and the mind-body problem. In Globus, G., Maxwell, G., and Savodinik, I., eds., Consciousness and the Brain. Plenum Press, pp. 205–266.
Wimsatt, W. C. (1976b). Reductive explanation, a functional account. In Hooker, C. A., Pearse, G., Michealos, A. C., and Evra, J. W. van, eds., Proceedings of the Meetings of the Philosophy of Science Association 1974. Reidel, pp. 671–710.
Winfree, A. T. (1984). The prehistory of the Belousov-Zhabotinsky oscillator. J. Chem. Educ., 61, 661–663.Google Scholar
Woese, C. (1967). The genetic code. New York: Harper & Row.
Woese, C. R. (1979). A proposal concerning the origin of life on the planet Earth. J. Mol. Evol., 13, 95–101.Google Scholar
Wolynes, P. G. (1997). As simple as can be? Nature Structural Biology, 11, 871–874.Google Scholar
Wong, J. T. (1975). A co-evolution theory of the genetic code. Proc. Natl. Acad. Sci. USA, 72, 1909–1912.Google Scholar
Wong, J. T. and Xue, H. (2002). Self-perfecting evolution of heteropolymer building blocks. In Fundamentals of Life, Editions scientifiques et medicales Elsevier SAS. Paris.
Wood, W. B. (1973). Genetic control of bacteriophage T4 morphogenesis. In Ruddle, F. J., ed., Genetic Mechanisms of Development. Academic Press, pp. 29–46.
Woodle, M. C. and Lasic, D. D. (1992). Biochim. Biophys. Acta, 1113, 171–199.
Yao, S., Ghosh, I., Zutshi, R., and Chmielewski, J. (1997). A pH-modulated self-replicating peptide. J. Am. Chem. Soc., 119, 10559–10560.Google Scholar
Yao, S., Ghosh, I., and Chmielewski, J. (1998). Selective amplification by auto- and cross-catalysis in a replicating peptide system. Nature, 396, 447–450.Google 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, 690–695.Google Scholar
Yarus, M. (2011). Getting past the RNA world: the initial Darwinian ancestor. Cold Spring Harb Perspect Biol., 3(4). doi:10.1101/cshperspect.a003590.Google Scholar
Yonath, A. (2010). Polar bears, antibiotics, and the evolving ribosome (Nobel Lecture). Angew Chem Int Ed Engl, 49, 4341–4354.Google Scholar
Yonath, A. (2012). Ribosomes: Ribozymes that Survived Evolution Pressures but Is Paralyzed by Tiny Antibiotics. In Carrondo, M. A. and Spadon, P., eds., NATO Science for Peace and Security Series A: Chemistry and Biology, Macromolecular Crystallography, pp. 195–208.
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, 689–696.Google Scholar
Yu, W., Sato, K., Wakabayashi, M., et al. (2001). Synthesis of functional protein in liposome. J. Biosc. Bioeng., 92, 590–593.Google Scholar
Yuen, G. U. and Knenvolden, K. A. (1973). Monocarboxylic acids in Murray and Murchison carbonaceous meteorites. Nature, 246, 301–302.Google 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, 43–47.Google Scholar
Zaher, H. S. and Unrau, P. J. (2007). Selection of an improved RNA polymerase ribozyme with superior extension and fidelity. RNA, 13, 1017–1026.Google 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, 325–337.Google 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.Google Scholar
Zeleny, M. (1977). Self-organization of living systems formal model of autopoiesis. Int. J. Gen. Syst., 4, 13–28.Google Scholar
Zelinski, W. S. and Orgel, L. E. (1987). Autocatalytic synthesis of a tetranucleotide analogue. Nature, 327, 346–347.Google Scholar
Zeng, F. W. and Zimmermann, S. C. (1997). Dendrimers in supramolecular chemistry: from molecular recognition to self-assembly. Chem. Rev., 97, 1681–1712.Google Scholar
Zepik, H. H., Bloechliger, E., and Luisi, P. L. (2001). A chemical model of homeostasis. Angew. Chem. Int. Ed. Engl., 40, 199–202.Google Scholar
Zhang, B. and Cech, T. R. (1998). Peptidyl-transferase ribozymes: trans reactions, structural characterization and ribosomal RNA-like features. Chem. Biol., 5, 539–553.Google Scholar
Zhao, M. and Bada, J. L. (1989). Extraterrestrial amino acids in cretaceous/tertiary boundary sediments at Stevns Klint, Denmark. Nature, 339, 463–465.Google Scholar
Zhao, Y., Liu, X., Wu, M., Tao, W., and Zhai, Z. (2000). In vitro nuclear reconstitution could be induced in a plant cell-free system. FEBS Letters, 480 (2–3): 208–212.Google Scholar
Zhu, J., Zhang, L., and Reszka, R. (1996). Liposome-mediated delivery of genes and oligonucleotides for the treatment of brain tumors. In Gregoriadis, G. and McCormack, B., eds., Targeting of Drugs: Strategies for Oligonucleide and Gene Delivery in Therapy. Plenum Press, pp. 169–187.
Zhu, J., Zhang, L., Hanisch, U. K., Felgner, P. L., and Reszka, R. (1996). In vivo gene therapy of experimental brain tumors by continuous administration of DNA-liposome complexes. Gene Therapy, 3, 472–476.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, 665–668.Google Scholar

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, Università degli Studi Roma Tre
  • Book: The Emergence of Life
  • Online publication: 05 September 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316135990.019
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, Università degli Studi Roma Tre
  • Book: The Emergence of Life
  • Online publication: 05 September 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316135990.019
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, Università degli Studi Roma Tre
  • Book: The Emergence of Life
  • Online publication: 05 September 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316135990.019
Available formats
×