Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-27T04:11:09.037Z Has data issue: false hasContentIssue false

11 - Tuning into the frequencies of life: a roar of static or a precise signal?

Published online by Cambridge University Press:  18 December 2009

Simon Conway Morris
Affiliation:
Professor of Evolutionary Palaeobiology, Earth Sciences Department, University of Cambridge
John D. Barrow
Affiliation:
University of Cambridge
Simon Conway Morris
Affiliation:
University of Cambridge
Stephen J. Freeland
Affiliation:
University of Maryland, Baltimore
Charles L. Harper, Jr
Affiliation:
John Templeton Foundation
Get access

Summary

Introduction

A glance at a bacterium, and a humpback whale, will reveal not only two immensely complex organisms, but also two very different life forms. Each is, in its respective way, constrained by a whole series of physical and chemical factors. One of the most obvious aspects is the fluid environment in which they live, albeit at scales that in being separated by about eight orders of magnitude are determinative of radically different behaviors. The bacterium's world is submillimetric, and accordingly it is dominated by constraints of viscosity. Motion involves the remarkable method of flagellar propulsion (the nearest thing to a wheel in biology; see, for example, Berg, 2003). When its flagella stop beating, the bacterium ceases its movement in a distance equivalent to the diameter of a hydrogen atom. The humpback whale, by contrast, occupies a liquid environment with which we are somewhat more familiar, although our swimming ability is feeble compared with the whale's oceanic travel range of thousands of kilometers. Fully aquatic, the humpback occupies a world that to us is both alien, with its complex system of echolocation, and familiar, with its ability to communicate – which includes singing.

Despite such wide divergences, the basic point of commonality is that both bacteria and whales live in environments where the physical controls imposed by the physico-chemical properties of water – be they viscosity or acoustic transmission – predetermine what is biologically possible. To this simple example could be added many other physical and chemical constraints.

Type
Chapter
Information
Fitness of the Cosmos for Life
Biochemistry and Fine-Tuning
, pp. 197 - 224
Publisher: Cambridge University Press
Print publication year: 2007

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

Allender, C. J.et al. (2003). Divergent selection during speciation of Lake Malawi cichlid fishes inferred from parallel radiations in nuptial coloration. Proceedings of the National Academy of Sciences, USA, 100, 14074–9.CrossRefGoogle ScholarPubMed
Axe, D. D. (2000). Extreme functional sensitivity to conservative amino acid changes on enzyme exteriors. Journal of Molecular Biology, 301, 585–96.CrossRefGoogle ScholarPubMed
Beeumen, J. J.et al. (1991). The primary structure of ruberythrin, a protein with inorganic pyrophosphate activity from Desulfovibrio vulgaris. Journal of Biological Chemistry, 266, 20645–53.Google Scholar
Beja, O.et al. (2001). Proteorhodopsin phototrophy in the ocean. Nature, 411, 786–9.CrossRefGoogle ScholarPubMed
Benner, S. A. (2000). Unite efforts and conquer mysteries of artificial genetics. Science, 290, 1506.CrossRefGoogle ScholarPubMed
Benner, S. A. (2002). The past as the key to the present: resurrection of ancient proteins from eosinophils. Proceedings of the National Academy of Sciences, USA, 99, 4760–1.CrossRefGoogle ScholarPubMed
Bennett, A. F. (2003). Experimental evolution and the Krogh principle: generating biological novelty for functional and genetic analyses. Physiological and Biochemical Zoology, 76, 1–11.CrossRefGoogle ScholarPubMed
Berg, H. C. (2003). The rotary motor of bacterial flagella. Annual Review of Biochemistry, 72, 19–54.CrossRefGoogle ScholarPubMed
Berg, J. S., Powell, B. C. and Cheney, R. E. (2001). A millennial myosin census. Molecular Biology of the Cell, 12, 780–94.CrossRefGoogle ScholarPubMed
Beuth, B., Niefind, K. and Schomburg, D. (2003). Crystal structure of creatinase from Pseudomonas putida: a novel fold and a case of convergent evolution. Journal of Molecular Biology, 332, 287–301.CrossRefGoogle Scholar
Boffelli, D., Cheng, J.-F. and Rubin, E. M. (2004). Convergent evolution in primates and an insectivore. Genomics, 83, 19–23.CrossRefGoogle ScholarPubMed
Bosch, T. C. G. and Khalturin, K. (2002). Patterning and cell differentiation in Hydra: novel genes and the limits of conservation. Canadian Journal of Zoology, 80, 1670–7.CrossRefGoogle Scholar
Brosset, A. (2003). Convergent and divergent evolution in rain-forest populations and communities of cyprinodontiform fishes (Aphyosemion and Rivulus) in Africa and South America. Canadian Journal of Zoology, 81, 1848–93.CrossRefGoogle Scholar
Bull, J. J.et al. (1997). Exceptional convergent evolution in a virus. Genetics, 147, 1497–1507.Google ScholarPubMed
Burmester, T. (2002). Origin and evolution of arthropod haemocyanins and related proteins. Journal of Comparative Physiology, B172, 95–107.Google Scholar
Charnock, S.-J.et al. (2002). Convergent evolution sheds light on the anti-β-elimination mechanism common to family 1 and 10 polysaccharide lyases. Proceedings of the National Academy of Science, USA, 99, 12067–72.CrossRefGoogle ScholarPubMed
Cheng, Z.et al. (2003). Highly divergent methyltransferases catalyze a conserved reaction in tocophenol and plastoquinone synthesis in cyanobacteria and photosynthetic eukaryotes. Plant Cell, 15, 2343–56.CrossRefGoogle Scholar
Coates, J. C. (2003). Armadillo repeat proteins: beyond the animal kingdom. Trends in Cell Biology, 13, 463–71.CrossRefGoogle ScholarPubMed
Cody, G. D.et al. (2000). Primordial carbonylated iron-sulphur compounds and the synthesis of pyruvate. Science, 289, 1337–40.CrossRefGoogle Scholar
Conant, G. C. and Wagner, A. (2003). Convergent evolution of gene circuits. Nature Genetics, 34, 264–6.CrossRefGoogle ScholarPubMed
Conway, Morris S. (2003.) Life's Solution: Inevitable Humans in a Lonely Universe. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Cooper, T. F., Rozen, D. E. and Lenski, R. E. (2003). Parallel changes in gene expression after 20,000 generations of evolution in Escherichia coli. Proceedings of the National Academy of Sciences, USA, 100, 1072–7.CrossRefGoogle Scholar
Coufal, D. E.et al. (2000). Sequencing and analysis of the Methylococcus capsulatus (Bath) soluble methane monooxygenase genes. European Journal of Biochemistry, 267, 2174–85.CrossRefGoogle Scholar
Crespi, B. J. (2001). The evolution of social behavior in microorganisms. Trends in Ecology and Evolution, 16, 178–83.CrossRefGoogle ScholarPubMed
Duve, C. (1995). Vital Dust: Life as a Cosmic Imperative. New York, NY: Basic Books (HarperCollins).Google Scholar
Dennison, K. L. and Spalding, E. P. (2000). Glutamate gated Ca2+ fluxes in Arabidopsis. Plant Physiology, 124, 1511–14.CrossRefGoogle Scholar
Denton, M. J., Marshall, C. J. and Legge, M. (2002). The protein folds as Platonic forms: new support for the pre-Darwinian conception of evolution by natural law. Journal of Theoretical Biology, 219, 325–42.CrossRefGoogle ScholarPubMed
Dose, K. (1988). The origin of life: more questions than answers. Interdisciplinary Science Reviews, 13, 348–56.CrossRefGoogle Scholar
Dupuy, F.et al. (2002). α1,4-fucosyltransferase activity: a significant function in the primate lineage has appeared twice independently. Molecular Biology and Evolution, 19, 815–24.CrossRefGoogle ScholarPubMed
Eizirik, E.et al. (2003). Molecular genetics and evolution of melanism in the cat family. Current Biology, 13, 448–53.CrossRefGoogle ScholarPubMed
Eschenmoser, A. (1999). Chemical etiology of nucleic acid structure. Science, 284, 2118–24.CrossRefGoogle ScholarPubMed
Fabrizio, J. J., Boyle, M. and DiNardo, S. (2003). A somatic role for eyes absent (eya) and sine oculis (so) in Drosophila spermatocyte development. Developmental Biology, 258, 117–28.CrossRefGoogle Scholar
Field, J.et al. (2000). Insurance-based advantage to helpers in a tropical hover wasp. Nature, 404, 869–71.CrossRefGoogle Scholar
Finnerty, J. R.et al. (2004). Origins of bilateral symmetry: Hox and Dpp expression in a sea anemone. Science, 304, 1335–7.CrossRefGoogle Scholar
Firn, E. (2004). Plant intelligence: an alternative point of view. Annals of Botany, 93, 345–51.CrossRefGoogle ScholarPubMed
Fukami, H.et al. (2004). Conventional taxonomy obscures deep divergence between Pacific and Atlantic corals. Nature, 427, 832–5.CrossRefGoogle ScholarPubMed
Gaucher, E. A.et al. (2003). Inferring the palaeoenvironment of ancient bacteria on the basis of resurrected proteins. Nature, 425, 285–8.CrossRefGoogle ScholarPubMed
Gillespie, R. (2004). Community assembly through adaptive radiation in Hawaiian spiders. Science, 303, 356–9.CrossRefGoogle ScholarPubMed
Gompel, N. and Carroll, S. B. (2003). Genetic mechanisms and constraints governing the evolution of correlated traits in drosophilid flies. Nature, 424, 931–5.CrossRefGoogle ScholarPubMed
Grammer, K.et al. (2003). Darwinian aesthetics: sexual selection and the biology of beauty. Biological Reviews, 78, 385–407.CrossRefGoogle ScholarPubMed
Grande, C.et al. (2004). Molecular phylogeny of Euthyneura (Mollusca: Gastropoda). Molecular Biology and Evolution, 21, 303–13.CrossRefGoogle Scholar
Gregory, T. R. (2004). Macroevolution, hierarchy theory, and the C-value enigma. Paleobiology, 3, 179–202.2.0.CO;2>CrossRefGoogle Scholar
Hamburger, Z. A.et al. (1999). Crystal structure of invasin: a bacterial integrin-binding protein. Science, 286, 291–5.CrossRefGoogle ScholarPubMed
Heanue, T. A.et al. (1999). Synergistic regulation of vertebrate muscle development by Dach2, Eya2, and Six1, homologs of genes required for Drosophila eye formation. Genes and Development, 13, 3231–43.CrossRefGoogle ScholarPubMed
Hedenström, A. (2004). A general law for animal locomotion. Trends in Ecology and Evolution, 19, 217–19.CrossRefGoogle ScholarPubMed
Henderson, L. J. (1913). The Fitness of the Environment: An Inquiry into the Biological Significance of the Properties of Matter. New York, NY: Macmillan. Repr. (1958) Boston, MA: Beacon Press; (1970) Gloucester, MA: Peter Smith.Google Scholar
Herrmann, K. M., Schultz, J. and Hermodson, M. A. (1980). Sequence homology between the tyrosine-sensitive 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase from Escherichia coli and hemerythrin from Sipunculida. Journal of Biological Chemistry, 255, 7079–81.Google ScholarPubMed
Hibbett, D. S., Grimaldi, D. and Donoghue, M. J. (1997a). Fossil mushrooms from Miocene and Cretaceous ambers and the evolution of the Homobasidiomycetes. American Journal of Botany, 84, 981–91.CrossRefGoogle Scholar
Hibbett, D. S.et al. (1997b). Evolution of gilled mushrooms and puffballs inferred from ribosomal DNA sequences. Proceedings of the National Academy of Sciences, USA, 94, 12002–6.CrossRefGoogle Scholar
Hillebrand, H. (2004). On the generality of the latitudinal diversity gradient. American Naturalist, 163, 192–211.CrossRefGoogle ScholarPubMed
Hoekstra, H. E. and Price, T. (2004). Parallel evolution is in the genes. Science, 303, 1779–81.CrossRefGoogle ScholarPubMed
Hofman, M. A. (2001). Brain evolution in hominids: are we at the end of the road? In Evolutionary Anatomy of the Primate Cerebral Cortex, ed. Falk, D. and Gibson, K. R.. Cambridge, UK: Cambridge University Press, pp. 113–27.CrossRefGoogle Scholar
Horgan, J. (1997). The End of Science: Facing the Limits of Knowledge in the Twilight of the Scientific Age. London: Little, Brown.Google Scholar
Hufford, L. (1997). The roles of ontogenetic evolution in the origins of floral homoplasies. International Journal of Plant Sciences, 158 (suppl. 6), 565–80.CrossRefGoogle Scholar
Immesberger, A. and Burmester, T. (2004). Putative phenoloxidases in the tunicate Ciona intestinalis and the origin of the arthropod hemocyanin superfamily. Journal of Comparative Physiology, B174, 169–80.CrossRefGoogle ScholarPubMed
Johns, G. C. and Somero, G. N. (2004). Evolutionary convergence in adaptation of proteins to temperature: A4-lactate dehydrogenase of Pacific damselfishes (Chromis spp.). Molecular Biology and Evolution, 21, 314–20.CrossRefGoogle Scholar
Jones, N. A.et al. (1998). The Drosophila Pax gene eye gone is required for embryonic salivary duct development. Development, 125, 4163–74.Google ScholarPubMed
Joshi, J. G. and Sullivan, B. (1973). Isolation and preliminary characterization of hemerythrin from Lingula unguis. Comparative Biochemistry and Physiology, B44, 857–67.Google Scholar
Kassam, D. D.et al. (2003). Morphometric analysis on ecomorphologically equivalent cichlid species from Lake Malawi and Tanganyika. Journal of Zoology, London, 260, 153–7.CrossRefGoogle Scholar
Knoll, A. H. and Bambach, R. K. (2000). Directionality in the history of life: diffusion from the left wall or repeated scaling of the right?Paleobiology, 26 (suppl.: Deep time, paleobiology's perspective), 1–14.CrossRefGoogle Scholar
Knouft, J. H. (2003). Convergence, divergence, and the effect of congeners on bodysize ratio in stream fishes. Evolution, 57, 2374–82.CrossRefGoogle Scholar
Koblmüller, S., Salzburger, W. and Sturmbauer, C. (2004). Evolutionary relationships in the sand-dwelling cichlid lineage of Lake Tanganyika suggests multiple colonization of rocky habitats and convergent origin of biparental mouthbrooding. Journal of Molecular Evolution, 58, 79–96.CrossRefGoogle Scholar
Kozlowski, J. and Konarzewski, M. (2004). Is West, Brown and Enquist's model of allometric scaling mathematically correct and biologically relevant?Functional Ecology, 18, 283–9.CrossRefGoogle Scholar
Kryukov, G. V.et al. (2002). Selenoprotein R is a zinc-containing stereo-specific methionine sulphoxide reductase. Proceedings of the National Academy of Sciences, USA, 99, 4245–50.CrossRefGoogle Scholar
Kurtz, D. M. (1999). Oxygen-carrying proteins: three solutions to a common problem. Essays in Biochemistry, 34, 85–100.CrossRefGoogle ScholarPubMed
Landi, M.et al. (2003). Low relatedness and frequent queen turnover in the stenogastrine wasp Eastenogaster fraterna favor the life insurance over the haplodiploid hypothesis for the origin of eusociality. Insectes Sociaux, 50, 262–7.CrossRefGoogle Scholar
Lebrun, E.et al. (2003). Arsenite oxidase, an ancient bioenergetic enzyme. Molecular Biology and Evolution, 20, 686–93.CrossRefGoogle ScholarPubMed
Lenski, R. E.et al. (2003). The evolutionary origin of complex features. Nature, 423, 139–44.CrossRefGoogle ScholarPubMed
Ligrane, R.et al. (2002). Diversity in the distribution of polysaccharide and glycoprotein epitopes in the cell walls of bryophytes: new evidence for the multiple evolution of water-conducting cells. New Phytologist, 156, 491–508.CrossRefGoogle Scholar
Linden, P. F. and Turner, J. S. (2004). “Optimal” vortex rings and aquatic propulsion mechanisms. Proceedings of the Royal Society, B271, 647–53.CrossRefGoogle ScholarPubMed
Litvak, Y. and Selinger, Z. (2003). Bacterial mimics of eukaryotic GTPase-activating proteins (GAPs). Trends in Biochemical Sciences, 28, 628–31.CrossRefGoogle Scholar
Locascio, A.et al. (2002). Modularity and reshuffling of Snail and Slug expression during vertebrate evolution. Proceedings of the National Academy of Sciences, USA, 99, 16841–6.CrossRefGoogle ScholarPubMed
Losos, J. B.et al. (2003). Niche lability in the evolution of a Caribbean lizard community. Nature, 424, 542–5.CrossRefGoogle ScholarPubMed
Mack, R. M. (2003). Phylogenetic constraint, absent life forms, and preadapted alien plants: a prescription for biological invasions. International Journal of Plant Sciences, 164 (suppl. 3), S185–96.CrossRefGoogle Scholar
Malcolm, B. A.et al. (1990). Ancestral lysozymes reconstructed, neutrality tested, and thermostability linked to hydrocarbon packing. Nature, 345, 86–9.CrossRefGoogle ScholarPubMed
Marden, J. H. and Allen, L. R. (2002). Molecules, muscles, and machines: universal performance characteristics of motors. Proceedings of the National Academy of Sciences, USA, 99, 4161–6.CrossRefGoogle ScholarPubMed
Matthysse, A. G.et al. (2004). A functional cellulose synthase from ascidian epidermis. Proceedings of the National Academy of Sciences, USA, 101, 986–91.CrossRefGoogle ScholarPubMed
Meehan, T. J. and Martin, L. D. (2003). Extinction and re-evolution of similar adaptive types (ecomorphs) in Cenozoic North American ungulates and carnivores reflect van der Hammen's cycles. Naturwissenschaften, 90, 131–5.Google Scholar
Mundy, N. I.et al. (2004). Conserved genetic basis of a quantitative plumage trait involved in mate choice. Science, 303, 1870–3.CrossRefGoogle ScholarPubMed
Nakashima, K.et al. (2004). The evolutionary origin of animal cellulose synthase. Development, Genes and Evolution, 214, 81–8.CrossRefGoogle ScholarPubMed
Nieder, A. and Miller, E. K. (2003). Coding of cognitive magnitude: compressed scaling of numerical information in the primate prefrontal cortex. Neuron, 37, 149–57.CrossRefGoogle ScholarPubMed
Nielsen, M. G.et al. (2003). Evolutionary convergence in otx expression in the pentameral adult rudiment in direct-developing sea urchins. Development, Genes and Evolution, 213, 73–82.Google ScholarPubMed
Donnell, O' K.et al. (2001). Evolutionary relationships among mucoralean fungi (Zygomycota): evidence for family polyphyly on a large scale. Mycologia, 93, 286–96.CrossRefGoogle Scholar
Pace, N. R. (2001). The universal nature of biochemistry. Proceedings of the National Academy of Sciences, USA, 98, 805–8.CrossRefGoogle ScholarPubMed
Pang, K., Matus, D. Q. and Martindale, M. Q. (2004). The ancestral role of COE genes may have been in chemoreception: evidence from the development of the sea anemone, Nematostella vectensis (Phylum Cnidaria: Class Anthozoa). Development, Genes and Evolution, 214, 134–8.CrossRefGoogle Scholar
Piatigorsky, J. (1992). Lens crystallins. Innovation associated with changes in gene regulation. Journal of Biological Chemistry, 267, 4277–80.Google ScholarPubMed
Primus, A. and Freeman, G. (2004). The cnidarian and the canon: the role of Wnt/β-catenin signalling in the evolution of metazoan embryos. BioEssays, 26, 474–8.Google Scholar
Queller, D. C. and Strassman, J. E. (1998). Kin selection and social insects. BioScience 48, 65–175.CrossRefGoogle Scholar
Rainey, P. B. and Rainey, K. (2003). Evolution of cooperation and conflict in experimental bacterial populations. Nature, 425, 72–4.CrossRefGoogle ScholarPubMed
Ranker, T. A.et al. (2004). Phylogeny and evolution of grammitid ferns (Grammitidaceae): a case of rampant morphological homoplasy. Taxon, 53, 415–28.CrossRefGoogle Scholar
Reeves, P. A. and Olmstead, R. G. (2003). Evolution of the TCP gene family in Asteridae: cladistic network approaches to understanding regulatory gene family diversification and its impact on morphological evolution. Molecular Biology and Evolution, 20, 1997–2009.CrossRefGoogle ScholarPubMed
Relaix, F. and Buckingham, M. (1999). From insect eye to vertebrate muscle: redeployment of a regulatory network. Genes and Development, 13, 3171–8.CrossRefGoogle ScholarPubMed
Rouse, G. W. and Pleijel, F. (2001). Polychaetes. Oxford: Oxford University Press.Google Scholar
Rudel, D. and Sommer, R. J. (2003). The evolution of developmental mechanisms. Developmental Biology, 264, 15–37.CrossRefGoogle ScholarPubMed
Savage, V. M.et al. (2004). The predominance of quarter-power scaling in biology. Functional Ecology, 18, 257–82.CrossRefGoogle Scholar
Schöning, K.-U.et al. (2000). Chemical etiology of nucleic acid structure: the α-threofuranosyl-(3′-2′) oligonucleotide system. Science, 290, 1347–51.CrossRefGoogle Scholar
Steel, J. B.et al. (2004). Are bryophyte communities different from higher-plant communities? Abundance relations. Oikos, 104, 479–86.CrossRefGoogle Scholar
Stenkamp, R. E. (1994). Dioxygen and hemerythrin. Chemical Reviews, 94, 715–26.CrossRefGoogle Scholar
Stiller, J. W., Reel, D. C. and Johnson, J. C. (2003). A single origin of plastids revisited: convergent evolution in organellar genome content. Journal of Phycology, 39, 95–105.CrossRefGoogle Scholar
Sucena, E.et al. (2003). Regulatory evolution of shavenbaby/ovo underlies multiple cases of morphological parallelism. Nature, 424, 935–8.CrossRefGoogle ScholarPubMed
Tabin, C. J., Carroll, S. B. and Panganiban, G. (1999). Out on a limb: parallels in vertebrate and invertebrate limb patterning and the origin of appendages. American Zoologist, 39, 650–63.CrossRefGoogle Scholar
Takagi, T. and Cox, J. A. (1991). Primary structure of myohemerythrin from the annelid Nereis diversicolor. FEBS Letters, 285, 25–7.CrossRefGoogle ScholarPubMed
Taylor, G. K., Nudds, R. L. and Thomas, A. L. R. (2003). Flying and swimming animals cruise at a Strouhal number tuned for high power efficiency. Nature, 425, 707–11.CrossRefGoogle Scholar
Theissen, G. and Becker, A. (2004). Gymnosperm orthologues of Class B floral homeotic genes and their impact on understanding flower origin. Critical Reviews in Plant Sciences, 23, 129–48.CrossRefGoogle Scholar
Trewavas, A. (2003). Aspects of plant intelligence. Annals of Botany, 92, 1–20.CrossRefGoogle ScholarPubMed
Trewavas, A. (2004). Aspects of plant intelligence: an answer to Firn. Annals of Botany, 93, 353–7.CrossRefGoogle ScholarPubMed
Velicer, G. J. and Yu, Y.-T N. (2003). Evolution and novel cooperative swarming in the bacterium Myxococcus xanthus. Nature, 425, 75–8.CrossRefGoogle ScholarPubMed
Volbeda, A. and Hol, W. G. J. (1989). Pseudo 2-fold symmetry in the copper-binding domain of arthropodan haemocyanins: possible implications for the evolution of oxygen transport proteins. Journal of Molecular Biology, 206, 531–46.CrossRefGoogle ScholarPubMed
Wagner, P. J. (2000). Exhaustion of morphologic character states among fossil taxa. Evolution, 54, 365–86.CrossRefGoogle ScholarPubMed
Wald, G. (1974). Fitness in the universe: choices and necessities. Origins of Life and Evolution of Biospheres, 5, 7–27.CrossRefGoogle ScholarPubMed
Weber, A. L. and Miller, S. L. (1981). Reasons for the occurrence of the twenty coded protein amino acids. Journal of Molecular Evolution, 17, 273–84.CrossRefGoogle ScholarPubMed
Weber, R. E. (1978). Respiratory pigments. In Physiology of Annelids, ed. Mill, P. J.. London: Academic Press, pp. 393–446.Google Scholar
Wessler, I., Kirkpatrick, C. J. and Racke, K. (1999). The cholinergic “pitfall”: acetylcholine, a universal cell molecule in biological systems, including humans. Clinical and Experimental Pharmacology and Physiology, 26, 198–205.CrossRefGoogle ScholarPubMed
West, G. B., Brown, J. H. and Enquist, B. J. (1999). The fourth dimension of life: fractal geometry and allometric scaling of organisms. Science, 284, 1677–9.CrossRefGoogle ScholarPubMed
West, G. B., Brown, J. H. and Enquist, B. J. (2001). A general model for ontogenetic growth. Nature, 413, 628–31.CrossRefGoogle ScholarPubMed
Westheimer, F. H. (1987). Why nature chose phosphates. Science, 235, 1173–8.CrossRefGoogle ScholarPubMed
Williams, R. J. P. and Fraústo da, Silva J. J. R. (2003). Evolution was chemically constrained. Journal of Theoretical Biology, 220, 323–43.CrossRefGoogle ScholarPubMed
Wittkopp, P. J.et al. (2003). Drosophila pigmentation evolution: divergent genotypes underlying convergent phenotypes. Proceedings of the National Academy of Sciences, USA, 100, 1808–13.CrossRefGoogle ScholarPubMed
Wolstencroft, R. D. and Raven, J. A. (2002). Photosynthesis: likelihood of occurrence and possibility of detection on Earth-like planets. Icarus, 157, 535–48.CrossRefGoogle Scholar
Wray, G. A. (2002). Do convergent developmental mechanisms underlie convergent phenotypes?Brain, Behavior and Evolution, 59, 327–36.CrossRefGoogle ScholarPubMed
Wray, G. A. and Lowe, C. J. (2000). Developmental regulatory genes and echinoderm evolution. Systematic Biology, 49, 28–51.CrossRefGoogle ScholarPubMed
Yoon, H. S. and Baum, D. A. (2004). Transgenic study of parallelism in plant morphological evolution. Proceedings of the National Academy of Sciences, USA, 101, 6524–9.CrossRefGoogle ScholarPubMed
Zakon, H. H. (2002). Convergent evolution on the molecular level. Brain, Behavior and Evolution, 59, 250–61.CrossRefGoogle ScholarPubMed
Zhang, J.-Z. and Rosenberg, H. F. (2002). Complementary advantageous substitutions in the evolution of an antiviral RNase of higher primates. Proceedings of the National Academy of Sciences, USA, 99, 5486–91.CrossRefGoogle ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

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.

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.

Available formats
×