Skip to main content

Part of the book series: NATO ASI Series ((NSSA,volume 259))

Abstract

“Which came first, the chicken or the egg?” For all its innocence, this is a question which, as this paper will show, obliges one to think deeply about the terms in which one ultimately seeks to explain biological systems. “A hen is only the egg’s way of making another egg”, in Butler’s (1878)1 apt phrase, sums up the view that organisms can most simply be explained as the results of causes laid out in the egg or, in contemporary terms, as the results of their inherited sets of genes. Today’s biologist might well extend the metaphor and say that “organisms are only expressions of a sequence of the four nucleotides of DNA”. This eminently successful way of explaining biological systems is examined in the first part of the present paper. Yet there remains a nagging worry about this viewpoint. Can this really be the whole story? In the second section, I attempt to identify the possible missing element and consider whether the other attributes of organisms might be more than mere genetic epiphenomena. In the third section I suggest how, due to the continuing impact of now outdated methodological assumptions, we face a choice of methods which greatly affects our ability to tackle the real problem of biological organization.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Preview

Unable to display preview. Download preview PDF.

Unable to display preview. Download preview PDF.

Notes and References

  1. Butlers. S. 1878, Life and Habit. London: Trubner.

    Google Scholar 

  2. Preformation dominated eighteenth century thinking about the origins of biological systems (Hartsoeker’s drawing dates from 1694), chiefly due to the overturning of the notion of spontaneous generation, on which the earlier idea of “epigenesis” had depended (see

    Google Scholar 

  3. Needham, J., 1934, A History of Embryology, Cambridge: Cambridge University Press.

    Google Scholar 

  4. Bowler, P. J., 1975, The changing meaning of “evolution”, J. Hist. Ideas 36: 95–114.

    Article  PubMed  CAS  Google Scholar 

  5. The “provisional theory of pangenesis” in Darwin, C., 1868, The Variation of Animals and Plants under Domestication, London: Murray.

    Google Scholar 

  6. It is difficult for us now to put ourselves back in Darwin’s conceptual framework, in which heredity and embryonic development were in effect indistinguishable. The embryo as such was poorly understood, even descriptively. His contemporary, Haeckel, considered embryonic stages (seen as recapitulating phylogenetic stages) to be the actual manifestations and vehicles of heredity. Against this view, His argued that embryogenesis must be explained in terms of physical causes at the time of development.

    Google Scholar 

  7. Weismann, A., 1893, The Germ-Plasm. A Theory of Heredity, London: Walter Scott. His theory was originally proposed in 1885.

    Google Scholar 

  8. Weismann hypothesized a process of assignment of specific nuclear formative factors to an organized pattern in the egg cytoplasm. Hertwig and Wilson, whose writings vividly illustrate the struggle to clarify the domains of genetics and embryology at the time, recognized clearly that Weismann’s theory implied preformed egg organization and that it did not explain embryonic regulation (first shown by Driesch in 1891) or regeneration.

    Google Scholar 

  9. Hertwig, O., 1896, The Biological Problem of To-day, London: Heinemann; Wilson, E.B., 1896, The Cell in Development and Inheritance, London: MacMillan.

    Google Scholar 

  10. Thompson, D’A. W., 1917, On Growth and Form, Cambridge: Cambridge University Press.

    Google Scholar 

  11. D’Arcy Thompson’s influence is notoriously difficult to characterize, but is probably considerable (see.

    Google Scholar 

  12. Medawar, P.B. in Thompson, R.D’A., 1958, D’Arcy Wentworth Thompson. The Scholar-Naturalist, 1860–1948, London: Oxford University Press.

    Google Scholar 

  13. Like His, his aim was “The study of organic Form by methods which are the common-place of physical science. It is the skilled and learned mathematician who must ultimately deal with such problems” (Preface). He was unconcerned with genetics or with embryology as such. As theories of development or evolution, modern versions of his approach (e. g. approaches based on the concepts of allometry and heterochrony) are very limited: they fail, for example, to explain changes in number or differentiation of body parts (see.

    Google Scholar 

  14. Horder, T.J. 1989. Syllabus for an embryological synthesis. In Complex Organismal Functions; Integration and Evolution in Vertebrates. Wake, D.B. and Roth, G., eds, pp 315–348. Chichester: Wiley.

    Google Scholar 

  15. “Neo-Darwinism” or “the modern evolutionary synthesis” of the 1930s and 1940s was primarily aimed at bridging the divide between genetics and evolution. Gradualistic evolution and adaptation were shown to be compatible (through population genetics) with the discrete effects of genes and mutations. The aim was achieved at the cost of considerable abstraction and detachment from specific data, particularly concerning phylogenesis, the specific forms of organisms and the actual phenomenology of embryogenesis. Changes in adult morphologies were seen as the results of gradualistic “growth” changes, using concepts like allometry, in turn linked to genetics in concepts such as “rate genes” and statistical gene combinations (Wright’s “shifting balance”). For analysis of methodological limitations see;.

    Google Scholar 

  16. Mayr, E., 1959, Where are we? Cold Spring Harbor Symp. Quant. Biol. 24: 1–14.

    Article  Google Scholar 

  17. Horder (1989) (op cit).

    Google Scholar 

  18. Regeneration, metaplasia and embryonic regulation all demonstrate the retention (and potential expressibility) of the entire genome in most cells throughout the life cycle, later confirmed by nuclear transplantation.

    Google Scholar 

  19. Figure 14.8.

    Google Scholar 

  20. Boveri, T., 1910, Die Potenzen der Ascaris-Blastomeren bei abgeanderter Furchung: Zugleich ein Beitrag zur Frage qualitativ ungleicher Chromosomenteilung, Festschr. R. Hertwig, Jena. III; 133-214.

    Google Scholar 

  21. Harrison, R.G., 1945, Relations of symmetry in the developing embryo, Trans. Conn. Acad. Arts Sci. 36: 277–330.

    Google Scholar 

  22. Spemann, H., 1938, Embryonic Development and Induction, New Haven: Yale University Press.

    Google Scholar 

  23. Waddington, C.H., 1940, Organisers and Genes, Cambridge: Cambridge University Press.

    Google Scholar 

  24. Waddington, C.H., 1957, The Strategy of the Genes, London: Allen and Unwin.

    Google Scholar 

  25. Figure 14.9; Derived from.

    Google Scholar 

  26. Wolpert, L., 1974, “The Development of Pattern and Form in Animals”, Oxford Biology Reader No 51. Oxford: Oxford University Press.

    Google Scholar 

  27. Figure 14.10; For details see.

    Google Scholar 

  28. Lawrence, P.A., 1992, The Making of a Fly. The Genetics of Animal Design, Oxford: Blackwell Scientific Publications.

    Google Scholar 

  29. On the positional information model, cells must have morphogen receptors capable of discriminating levels to a degree equal to the number of morphological positions at which cells undergo alternative forms of differentiation. A particular difficulty for locus-specific control mechanisms concerns the differentiation of identical structures at multiple, different locations: on the positional information model, specificity would still be needed at each locus even if the same cell type differentiates (as acknowledged in the associated concept of “non-equivalence”). This independence of locus receptivity and differentiation makes it difficult to see how evolution could change pattern, since two, matching mutations serving the two functions would always be needed simultaneously.

    Google Scholar 

  30. Lewis, E.B., 1978, A gene complex controlling segmentation in Drosophila. Nature 276: 565–570.

    Article  PubMed  CAS  Google Scholar 

  31. “It has become almost a truism to assume that the position of a gene, either on a chromosome or else within a nucleus, is of critical importance to its functional capabilities”.

    Google Scholar 

  32. John, B. and Miklos, G., 1988, The Eukaryote Genome in Development and Evolution, London: Allen and Unwin, p193.

    Google Scholar 

  33. The concept of master genes — hypothesized as functioning to control and integrate groups of genes mediating pattern or mediating the differentiation of specific cell types (see

    Google Scholar 

  34. Davidson, E.H. and Britten, R.J., 1979, Regulation of gene expression: possible role of repetitive sequences. Science 204: 1052–1059.

    Article  PubMed  CAS  Google Scholar 

  35. MacIntyre, R.J., 1982, Regulatory genes and adaptation: past, present, and future, Evol. Biol., 15: 247–285.

    Article  CAS  Google Scholar 

  36. — carries preformationist implications. Such overall control elements may be unnecessary given that control might be explained as the result of a cascade, whereby the products of one structural gene are used as signals for another, and so on in a chain reaction.

    Google Scholar 

  37. Although maternally imposed patterning in the egg cytoplasm is well established, this mechanism has occupied a diminishing place in developmental theories. Adequate testing increasingly reveals other mechanisms (e. g. see.

    Google Scholar 

  38. Lambie, E.J. and Kimble, J., 1991, Genetic control of cell interactions in nematode development. Ann. Rev. Genet. 25: 411–436). Most maternal factors are involved in forming the egg cell as such and their role in pattern control must be restricted if they are not to limit zygotic variation and rates of possible future evolution. Compared to other patterning mechanisms, their role is always very limited and is minimal in vertebrates.

    Article  PubMed  CAS  Google Scholar 

  39. Examples of functional modulation of morphology are reviewed in.

    Google Scholar 

  40. Goss, R.J., 1978, The Physiology of Growth, New York: Academic Press.

    Google Scholar 

  41. Horder, T.J., 1983, Embryological bases of evolution. In Development and Evolution, Goodwin, B.C., Holder, N. and Wylie, C.C, eds, pp 315–352. Cambridge: Cambridge University Press.

    Google Scholar 

  42. Holmes, S.J., 1948, Organic Form and related Biological Problems, Berkeley: University of California Press.

    Google Scholar 

  43. Why such evidence has not generally featured in theories of development is unclear; apart from the Lamarckian overtones and the difficulties of defining the underlying genetic and cellular mechanisms (in common with many developmental phenomena), 20 there is perhaps a tendency to think of “environmental factors” as external to the organism: i. e. to disregard “internal environmental” factors.

    Google Scholar 

  44. Darnell, J., Lodish, H. and Baltimore, D., 1990, Molecular Cell Biology, New York: Scientific American Books.

    Google Scholar 

  45. Three types of cause, commonly combined in biological phenomena, can loosely be distinguished; the source of the specificity of the factor or function under consideration (an “instructive” cause); the conditions needed to trigger or regulate events (“elective”); and other preconditions necessary for events, but not directly contributing to them (“facultative”, “permissive”).

    Google Scholar 

  46. There are many contending definitions of the term “genetic”.

    Google Scholar 

  47. Kitcher, P., 1982, Genes Brit. J. Phil. Sci. 33; 337–359.

    Article  Google Scholar 

  48. Even the nearest to an ultimate definition — the description of a specific DNA sequence — involves ambiguities (e. g. whether to include variations between individuals, introns, initiating and control regions). The concept of “genetic” factors is not just difficult to define: it is ultimately meaningless, because it cannot be separated from “environmental” factors. The two types of consideration will always interact, because no genetic effects can occur unless they are expressed and expression depends on factors extrinsic to the DNA sequence itself. What is more, there is no complete way of determining the extent of involvement of either consideration, given the complexity of their interactions and the practical impossibility of testing all possible environmental variants. Despite the importance and frequency of the appeals to a supposed separability (as in “nature/nurture” discussions), these points are rarely made and the literature on the subject is extraordinarily restricted. For the best available attempts at clarification see.

    Google Scholar 

  49. Oyama, S., 1985, The Ontogeny of Information. Developmental Systems and Evolution, Cambridge: Cambridge University Press.

    Google Scholar 

  50. Rose, S., Kamin, L.J. and Lewontin, R.C., 1984, Not in Our Genes, Harmondsworth: Penguin Books.

    Google Scholar 

  51. For an important critique of geneticism see.

    Google Scholar 

  52. Tauber, A.I. and Sarkar, S., 1992, The human genome project: has blind reductionism gone too far? Perspect. Biol. Med. 35: 220–235.

    PubMed  CAS  Google Scholar 

  53. The gene concept has always referred primarily to a structural entity. In practice “identifying the gene for a function” means localizing an involved chromosomal site, or sequencing it or its products. These operations are usually merely handles for further practical procedures: the specificity of the products and their functions are usually in themselves of secondary importance. Compared to their structure, the functions of genes, via their products, are difficult to investigate and can never be delimited with certainty. Despite the problems, the search for genetic explanations continues apace.

    Google Scholar 

  54. Owen, M. and McGuffin, P., 1992, The molecular genetics of schizophrenia, Brit. Med. J. 305: 664–665.

    Article  PubMed  CAS  Google Scholar 

  55. Alper, J.S. and Natowicz, M.R., 1992, The allure of genetic explanations, Brit. Med. J. 305: 666.

    Article  PubMed  CAS  Google Scholar 

  56. Figure 14.11; adapted from.

    Google Scholar 

  57. Patten, B.M., 1946, Human Embryology, London: Churchill.

    Google Scholar 

  58. The phenomenon of induction is massively documented and Figure 14.11 summarises many of the known cases. See also.

    Google Scholar 

  59. Nakamura, O. and Toivonen, S., 1978, Organizer — a milestone of a half-century from Spemann, Elsevier: Amsterdam.

    Google Scholar 

  60. Regulation, coordination and precision (potentially to the level of individual cells) of adult pattern (the main reasons for hypothesizing gradients) are explicable by the developmental cascade, as are varieties of distributions and combinations of differentiated cell types. Induction has been a focus for much dispute, largely due to methodological problems (e. g. multi-causality due to double assurance, involvement of competence, reciprocal interactions between inductor and induced tissues) inherent in the analysis of any higher level embryological phenomena. For reviews of these problems and the role of morphogenetic movements and receptivity to inductors see.

    Google Scholar 

  61. Horder (1983) (op cit).

    Google Scholar 

  62. Horder, T.J., 1976, Pattern formation in animal development. In The Developmental Biology of Plants and Animals, Graham, C.F. and Wareing, P.F., eds, pp 169–197. Oxford: Blackwell Scientific Publications.

    Google Scholar 

  63. Induction and positional information are opposing concepts. In induction the control of one cell’s fate is locally determined; only “position” in relation to neighbouring cells is relevant and it is unnecessary that a cell has information about its position in the embryo as a whole. For further discussion see; Hinchliffe and Horder (this volume).

    Google Scholar 

  64. The genetic analysis of Drosophila development (Figure 14.10) illustrates some of the limitations of genetic methodology in general. It is essentially descriptive. Increasingly, it is becoming clear that a full explanation of development requires more than just the identification of genes or description of their products and their spatio-temporal patterns of expression; it also requires explanation of the distributions and interactions of these products, since it is these considerations which determine succeeding gene switches. It is likely that any one switching factor (e. g. the bicoid gradient) is only responsible for a limited number (around three) of spatially distinct patterns as expressed in the next round of gene switches. Pattern is therefore actually built up gradually (i. e. involving multiple interactive, dependent stages — the gene product of one stage having free, direct access to the promoters of other genes which are then selectively switched14 — so that final adult differentiation is not the result of direct one-to-one read out of initial position, e. g. as defined by the bicoid gradient; interactions include induction and morphogenesis.

    Google Scholar 

  65. Horder, 1983, op cit.

    Google Scholar 

  66. Early embryogenesis and mutations affecting segmentation in Drosophila may be poor models for other organisms; e. g. Drosophila embryos do not regulate (see.

    Google Scholar 

  67. Sander in Goodwin et al. (1983) (op cit).

    Google Scholar 

  68. and equivalent (“homeotic”) mutations are unknown in vertebrates.

    Google Scholar 

  69. Horder, 1976;1983, op cit).

    Google Scholar 

  70. See Horder (1989) (op cit).

    Google Scholar 

  71. which includes arguments against the notion of pattern (e. g. limb pattern) genes. There is a wide spectrum of possible relationships between genes and their ranges of effect morphologically; e. g. genes may be locus or organspecific, cell-type specific (with mutations often leading to patchy effects) or may affect metabolism (with mutations having diffuse effects). In all cases pleiotropy may be due to knock-on effects, secondary to the primary gene action. There are major limits to our ability to analyse the genetic programming of pattern; e. g. due to incomplete identification of genes or mutations and difficulties in separating component genes in situations of polygenetic control.

    Google Scholar 

  72. On the evolutionary flexibility of chromosomal organization see.

    Google Scholar 

  73. Berry, R.J., 1977, Inheritance and Natural History, London: Collins.

    Google Scholar 

  74. Wasserman, M., and Wasserman, F., 1992, Inversion polymorphism in island species of Drosophila, Evol. Biol. 26: 351–381.

    Article  Google Scholar 

  75. O’Brien, S.J., and Seuanez, H.N., 1988, Mammalian genome organization: an evolutionary view, Ann. Rev. Genetics 22: 323–351

    Article  Google Scholar 

  76. John and Miklos (1988) (op cit).

    Google Scholar 

  77. As an example of the dispersed mapping of protein molecules originating by duplication from one ancestral form, see.

    Google Scholar 

  78. Wilkie, T.M., Gilbert, D.J., Olsen. S.A., Chen, X-N., Amatruda, T.T., Korenberg, J.R., Trask., B.J., de Jong, P., Reed, R.R., Simon, M.I., Jenkins, N.A. and Copeland, N.G., 1992, Evolution of the mammalian G protein alpha-subunit multigene family, Nature Genetics, 1: 85–91.

    Article  PubMed  CAS  Google Scholar 

  79. Haemoglobin sub-units illustrate dispersed mapping of a single final molecule: see also.

    Google Scholar 

  80. Wissinger, B., Schuster, W., and Brennicke, A., 1991, Trans splicing in Oenotheran mitochondria: nad1 mRNAs are edited in exon and transsplicing Group II intron sequences, Cell 65: 473–482.

    Article  PubMed  CAS  Google Scholar 

  81. Bacteria may not be good models here; they typically show integrated polycistronic genome control and very stable gene sequences in evolution.

    Google Scholar 

  82. Riley, M. and Krawiec, S., 1987, Genome organization. In Escherichia coli and Salmonella typhimurium: Cellular and Molecular Biology, Neidhardt, F.C., Ingraham, J.L., Low, K.B., Magasanik, B., Schaechter, M. and Umbarger, H.E. eds. pp 967–981. Washington: American Society of Microbiology.

    Google Scholar 

  83. In general gene organization, in higher organisms, tends towards random order. Given appropriate control mechanisms14,22 this is entirely compatible with integrated function. Many special factors may explain instances where particular gene sequences show close relationships; e. g. recent evolution by gene duplication, “supergenes”, “position effects”, linkage disequilibrium, “gene complexes”. On the lack of effect of dissociating neighbouring genes see.

    Google Scholar 

  84. Struhl, G., 1984, Splitting the bithorax complex in Drosophila, Nature 308: 454–457.

    Article  Google Scholar 

  85. John and Miklos (1988) (op cit).

    Google Scholar 

  86. See Horder, T.J., 1991, Molecular biology and evolution: two perspectives. In Developmental Patterning of the Vertebrate Limb, Hinchliffe, J.R., Hurle, J.M. and Summerbell, D. eds, pp 423–438. New York: Plenum.

    Chapter  Google Scholar 

  87. Horder, T.J., 1993, Three glimpses of evolution. In Formation and Regeneration of Nerve Connections, Sharma, S.C. and Fawcett, J.W., eds. pp 222–238. Boston: Birkhauser.

    Google Scholar 

  88. Nei, M. and Koehn, R.K., 1983, Evolution of Genes and Proteins, Sunderland, MA: Sinauer.

    Google Scholar 

  89. Li, W-H. and Graur, D., 1991, Fundamentals of Molecular Evolution, Sunderland, MA: Sinauer.

    Google Scholar 

  90. “Causality” in the organism may sometimes come to differ widely from the causal chain of events in evolution; e. g. in the case of DNA as the “cause” of RNA and in turn proteins, it seems likely that the evolution of RNA preceded that of DNA (see.

    Google Scholar 

  91. Darnell et al. (1990) (op cit).

    Google Scholar 

  92. Chapter 26). However, other characters (such as the DNA genetic code itself) are so constant and universal among organisms that we can confidently infer their early origin and subsequent evolutionary inflexibility. Sequences of developmental events in extant organisms often provide direct evidence on which evolutionary inferences can be based and sometimes “recapitulate” or retain the record of the evolutionary sequences of events.

    Google Scholar 

  93. Horder, 1993, op cit.

    Google Scholar 

  94. Recombination is evidently almost as basic a property of DNA sequences as the DNA code itself. Maintaining the “integration” of a genome tending towards random order becomes a matter, not only of the selective advantages, matching and compatibility of structural gene products, but also of the regulation of the units of DNA rearrangement.

    Google Scholar 

  95. Plasterk, R.H.A., 1992, Genetic switches: mechanisms and function, Trends in Genetics, 8: 403–406.

    PubMed  CAS  Google Scholar 

  96. and expression control, 11,14 by other, nonexpressed DNA components.

    Google Scholar 

  97. It is easy to think of natural selection as only operating directly on the adult; however, despite its protected circumstances, the embryo is indirectly under even greater selective pressures because multiple adult characters depend on each developmental step; hence the relative evolutionary stability of developmental processes (Horder, 1993, op cit).

    Google Scholar 

  98. The available terminology is inadequate;39 the term “epigenesis” comes closest to covering the concept we have attempted to characterize, but is unsatisfactory because of its connotations as the antithesis of preformation and of genetics.

    Google Scholar 

  99. Roll-Hansen, N., 1978, Drosophila genetics: a reductionist research program, J. Hist. Biol. 11: 159–210.

    Article  Google Scholar 

  100. On vitalism, physico-chemical reduction and the origin and use of “models”, see.

    Google Scholar 

  101. Hall, T.S., 1969, History of General Physiology, Chicago: University of Chicago Press.

    Google Scholar 

  102. Oppenheimer, J.M., 1967, Essays in the History of Embryology and Biology, Cambridge: M.I.T. Press.

    Google Scholar 

  103. Haraway, D.J., 1976, Crystals, Fabrics, and Fields. Metaphors of Organicism in Twentieth-Century Developmental Biology, New Haven: Yale University Press.

    Google Scholar 

  104. Oyama (1985) (op cit).

    Google Scholar 

  105. On the origin of embryological concepts, see.

    Google Scholar 

  106. Needham (1934) (op cit).

    Google Scholar 

  107. Holmes (1948) (op cit).

    Google Scholar 

  108. Oppenheimer (1967) (op cit).

    Google Scholar 

  109. Haraway (1976) (op cit).

    Google Scholar 

  110. Horder, T.J., Witkowski, J.A. and Wylie, C.C., 1985, A History of Embryology, Cambridge: Cambridge University Press.

    Google Scholar 

  111. Gilbert, S.F., 1991 A Conceptual History of Modern Embryology, New York: Plenum.

    Book  Google Scholar 

  112. In this area of biology — notable for its importation of ex-physical scientists, e. g. Turing, Schrodinger, Whitehead, Weiss, Wolpert — many concepts have been borrowed from the physical sciences; e.g. field, polarity, system, double assurance. On the coextensivity of thinking in physical and biological sciences see, for example.

    Google Scholar 

  113. Stebbing, L.S., 1937, Philosophy and the Physicists, Harmondsworth: Penguin.

    Google Scholar 

  114. Hull, D.L., 1974, Philosophy of Biological Science, Englewood Cliffs: Prentice-Hall.

    Google Scholar 

  115. Yoxen, E.J., 1979, Where does Schroedinger’s “What is Life?” belong in the history of molecular biology? Hist. Science, 17: 17–52.

    CAS  Google Scholar 

  116. As case histories, see.

    Google Scholar 

  117. Whyte, L.L., 1963, Focus and Diversions, London: Cresset Press.

    Google Scholar 

  118. Fischer, E.P. and Lipson, C., 1988, Thinking about Science. Max Delbruck and the Origin of Molecular Biology, New York: Norton.

    Google Scholar 

  119. “The purpose of this paper is to discuss a possible mechanism by which the genes of a zygote may determine the anatomical structure of the resulting organism. The theory does not make any new hypotheses; it merely suggests that certain well-known physical laws are sufficient to account for many of the facts”.

    Google Scholar 

  120. Turing A.M., 1952, The chemical basis of morphogenesis, Phil. Trans Roy. Soc. 237B: 37–72.

    Google Scholar 

  121. Problems of biological explanation often revolve around a conceptual dichotomy or antithesis, both elements of which very often turn out to be interrelated and equally relevant in the end (e. g. nature/nurture)19 or the bridging of the domains of evolution and genetics.

    Google Scholar 

  122. Horder, 1989, op cit.

    Google Scholar 

  123. Such disjunctions often originate in the way in which data are initially categorized and classified; e. g. the distinguishing of genetics and embryology (as described above), pre-functional and functional phases of development.

    Google Scholar 

  124. Holmes, 1948, op cit.

    Google Scholar 

  125. or morphogenesis from pattern formation and differentiation (see).

    Google Scholar 

  126. Waddington, 1957, op cit.

    Google Scholar 

  127. Wolpert, 1974, op cit.

    Google Scholar 

  128. Many biological concepts are used in a disjunctive or all-or-none manner (e. g. homology, species, phases of life cycle, genotype/phenotype, nature/nurture) leading to typological and saltationist thinking, which can often be seen to create artefacts which misrepresent the underlying continuities of biological processes.38.

    Google Scholar 

  129. It could be argued that in practice what has happened is that, caught between the far more coherent and well-developed fields of genetics and evolution theory, embryology has been conceptualized in conformity with assumptions and expectations derived from other fields.

    Google Scholar 

  130. On the theory of “reducibility of theories” and transitions between levels, see.

    Google Scholar 

  131. Nagel, E., 1961, The Structure of Science, London: Routledge and Kegan Paul.

    Google Scholar 

  132. It has become increasingly recognized that reduction cannot be fully realized in practice, even in the paradigm case of translating between Mendel’s laws and molecular genetics: the rules of translation cannot themselves be closely enough specified (see.

    Google Scholar 

  133. Rosenberg, A., 1985, The Structure of Biological Science, Cambridge: Cambridge University Press.

    Book  Google Scholar 

  134. Hull (1974) (op cit).

    Google Scholar 

  135. Hull, D.L., 1976, Informal aspects of theory reduction, Boston Studies in the Phil. of Science, 32: 653–670.

    Article  Google Scholar 

  136. Similar problems apply to axiomatization; see.

    Google Scholar 

  137. Ruse, M., 1975, Woodger on genetics. A critical evaluation, Acta Biotheoret. 24: 1–13.

    Article  CAS  Google Scholar 

  138. Given a large enough scale of view, biological functions and processes (with very few exceptions and in marked contrast to biological structures) can be seen as continuous; e. g. continuity of genome and cell organization across generations, morphogenesis, phases of the life cycle, growth, morphological integration. The discrete manifestations of evolution immediately available to us (e. g. the individuality of organisms, gaps between known species, the quantal nature of mutations) encourage disjunctive thinking.35 However, there must be limits to the size of unit steps of evolutionary change compatible with the survival of intermediate stages (particularly given the requirements for genome integration)11,28 and the concept of the species as a gene pool containing a reservoir of possible variants implies a diffuse, statistical basis for change: the nature of embryogenesis provides strong grounds for inferring a close continuity of morphologies during phylogenesis.

    Google Scholar 

  139. Horder, 1993, op cit.

    Google Scholar 

  140. The term “emergence” refers to the appearance of new (and, by implication, unpredictable) properties as the result of the combination of elements in a complex system.

    Google Scholar 

  141. Nagel, 1961, op cit.

    Google Scholar 

  142. As Nagel argues, whether new properties are “unpredicted” all depends on how well understood the contributing elements are in the first place: if the complexity of the system is adequately recognized the emergent properties become less surprising. A single complex molecule is often a sufficient and fully understandable explanation for high level physiological functions (e. g. haemoglobin). A type of thinking based on emergence underwrites many anti-reductionist claims about biological systems.

    Google Scholar 

  143. Simpson, G.G., 1963, Biology and the nature of science, Science 139: 81–88.

    Article  PubMed  CAS  Google Scholar 

  144. Polanyi, M. 1968, Life’s irreducible structure, Science 160: 1308–1312.

    Article  PubMed  CAS  Google Scholar 

  145. Weiss, P.A., 1971, Hierarchically Organized Systems in Theory and Practice, New York: Hafner.

    Google Scholar 

  146. On laws in biology, see.

    Google Scholar 

  147. Rosenberg (1985) (op cit).

    Google Scholar 

  148. Due to the opportunistic nature of evolution, it is inevitable that there will be few broad generalizations in biology and that most will eventually have exceptions. Only at the level of chemistry (i. e. molecular biology) will laws (in the usual sense of absolute rules) apply. High level concepts such as evolution are more like generalizations than laws.

    Google Scholar 

  149. Fundamental to what I have been saying is the importance of an awareness of the procedures involved in the use of data and of the concepts needed to link them. Embryology has been dogged by the inadequacies of its heritage of concepts and terminology (e. g. epigenesis, organicism, holism, emergence, etc): suspicions regarding their imprecision, abstractness or merely metaphorical character, and of the theories built on them, were often justified.

    Google Scholar 

  150. That the priorities of philosophers of biology should lag behind those of biologists is perhaps inevitable given the philosophers’ dependence on the state of already established scientific knowledge and their traditional affinities with the physical sciences.

    Google Scholar 

  151. Hull, D.L., 1969, What philosophy of biology is not, J. Hist Biol. 2: 241–268.

    Article  Google Scholar 

  152. Hull (1974) (op cit).

    Google Scholar 

  153. For a penetrating analysis of the requirements of scientific method in biology (as applied to psychology), see.

    Google Scholar 

  154. Kaplan, A., 1964, The Conduct of Inquiry, San Francisco: Chandler.

    Google Scholar 

  155. Nagel (1961, op cit).

    Google Scholar 

  156. defines “explanation” as “systematized knowledge”.

    Google Scholar 

  157. Kaplan (1964, op cit, p329).

    Google Scholar 

  158. as “concatenated description… each element of what is being described shines, as it were, with light reflected from all the others”.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 1993 Springer Science+Business Media New York

About this chapter

Cite this chapter

Horder, T.J. (1993). The Chicken and the Egg. In: Othmer, H.G., Maini, P.K., Murray, J.D. (eds) Experimental and Theoretical Advances in Biological Pattern Formation. NATO ASI Series, vol 259. Springer, Boston, MA. https://doi.org/10.1007/978-1-4615-2433-5_14

Download citation

  • DOI: https://doi.org/10.1007/978-1-4615-2433-5_14

  • Publisher Name: Springer, Boston, MA

  • Print ISBN: 978-1-4613-6033-9

  • Online ISBN: 978-1-4615-2433-5

  • eBook Packages: Springer Book Archive

Publish with us

Policies and ethics