Skip to main content
Log in

Epistemic, Evolutionary, and Physical Conditions for Biological Information

  • Original Paper
  • Published:
Biosemiotics Aims and scope Submit manuscript

Abstract

The necessary but not sufficient conditions for biological informational concepts like signs, symbols, memories, instructions, and messages are (1) an object or referent that the information is about, (2) a physical embodiment or vehicle that stands for what the information is about (the object), and (3) an interpreter or agent that separates the referent information from the vehicle’s material structure, and that establishes the stands-for relation. This separation is named the epistemic cut, and explaining clearly how the stands-for relation is realized is named the symbol-matter problem. (4) A necessary physical condition is that all informational vehicles are material boundary conditions or constraints acting on the lawful dynamics of local systems. It is useful to define a dependency hierarchy of information types: (1) syntactic information (i.e., communication theory), (2) heritable information acquired by variation and natural selection, (3) non-heritable learned or creative information, and (4) measured physical information in the context of natural laws. High information storage capacity is most reliably implemented by discrete linear sequences of non-dynamic vehicles, while the execution of information for control and construction is a non-holonomic dynamic process. The first epistemic cut occurs in self-replication. The first interpretation of base sequence information is by protein folding; the last interpretation of base sequence information is by natural selection. Evolution has evolved senses and nervous systems that acquire non-heritable information, and only very recently after billions of years, the competence for human language. Genetic and human languages are the only known complete general purpose languages. They have fundamental properties in common, but are entirely different in their acquisition, storage and interpretation.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Hoffmeyer (2008, p. 93) speaks of my subscribing to “the ontology of natural law.” Physics says nothing about the ontology of natural law besides the belief that some unknowable kind of reality exists outside our mental images and symbols. There are only epistemological and objectivity conditions. Objectivity requires natural laws to appear common to all conceivable observers. Hertz’s briefly stated epistemology is accepted by most physicists: “We form for ourselves images or symbols of external objects; and the form which we give them is such that the logically necessary consequents of the images in thought are always the images of the necessary natural consequents of the thing pictured. . . For our purpose it is not necessary that they [the images] should be in conformity with the things in any other respect whatever. As a matter of fact, we do not know, nor have we any means of knowing, whether our conception of things are in conformity with them in any other than this one fundamental respect” (Hertz 1984).

  2. Hoffmeyer refers to one paper (Pattee 1997) in which I used the linguist’s narrow definition of semiosis, as I have done in other papers. That was before I was aware of its wider use in biosemiotics. I have never thought of information as limited to discrete symbols. For example I said (Pattee 1989), “The structure of the folded string has both a shape and an activity, neither of which could be called arbitrary. In fact, the folded enzyme can be called both iconic in its binding site and mimetic in its catalytic activity. The folding process as well as the substrate binding and catalysis are no longer local, sequential, or context-free, but highly parallel and global. Yet physically the folded enzyme contains the same molecules and obeys the same laws as the string of symbol vehicles. What has happened is that those aspects of structure that are essential for the activities as symbol vehicles have been suppressed, while other aspects of the symbol vehicle structure that are excluded from explicit symbolic activity, have been brought into play. Furthermore, these two aspects appear as complementary in the sense that one aspect cannot function in both roles simultaneously. A structure cannot be both arbitrary and iconic, or rate-independent and rate-dependent, or discrete and continuous at the same time. However, the same structure can possess any of these aspects depending on its context, i.e., the level of organization where the particular aspect is brought into play” (p. 271). I give a detailed discussion of the context-dependence of discrete and continuous descriptions in Pattee (1974).

  3. Here is my conclusion of The Physical Basis of Coding and Reliability in Biological Evolution presented at the 1966 Bellagio meeting (Pattee 1968, p. 89). “At the evolutionary level this concept of a symbolic genetic description and its code structures must be broadened to a larger system that includes not only the description of the system itself but also a description or a 'theory' of the environment. In the evolutionary context the phenotype itself now plays the role of a composite measuring device that tests the descriptive theory through its interactions with the real environment. In this language we must also expand the concept of reliability to include the overall predictive value of this description-code or theory-measurement system. I believe it is then reasonable to associate this overall predictive value with what is called the 'measure of fitness' in evolutionary theory. Finally, at the level of nervous activity in the processes of memory and intellectual theory making, we are again searching for more elegant code structures which allow the maximum predictive reliability over the widest domain, but which can be generated from relatively short symbolic descriptions. Perhaps we could even say that the characteristic sign of biological activity at all levels is the existence of efficient and reliable codes”

  4. Wigner replied to my argument: “I believe I understand your arguments in this regard and concur with you. The reason for my arguing on the basis of consciousness was indeed that in this case I could adduce evidence for the incompleteness, whereas I could not do this at a lower level” (Pattee 1972, Note 1). I am not sure that Léon Rosenfeld was complelely convinced. After a year of correspondence he cautiously concluded, “I did not meet any statement with which I would disagree.” However, he emphasized that my discussion, “was irrelevant to the problem of the actual measuring process” with which I agree because I have not addressed the measurement problem itself.

  5. Classically, a non-holonomic constraint is a relation between the variables which limits more degrees of freedom of the dynamic motion than the static degrees of freedom that are necessary to specify the initial conditions (e.g. Sommerfeld 1952) In other words, the number of constrained dimensions, f–r, of the actual trajectory space of a system with non-holonomic constraints is smaller than the number of dimensions, f, necessary for a complete state description, by the number of non-holonomic relations, r. The constrained trajectory space, Rf–r is a subspace of the state space, Sf, but an external perturbation or mutation in the constraint can produce a different trajectory subspace, R′f–r. It is only by coupling such non-holonomic systems together that we can build machines. Computers are an extreme case of a maximally constrained machine with only enough dynamics to drive one state to the next. As I discussed, living systems are minimally constrained by genetic information, because they rely heavily on law-based dynamics. For a discussion of non-holonomic constraints in quantum theory see Eden (1951).

  6. Watch a folding simulation video at http://www.youtube.com/watch?v=gFcp2Xpd29I More information on folding at http://folding.stanford.edu/English/Science

  7. Peirce’s discussion of his logical graphs illustrates one of his common lapses from clear logic to vague metaphysics. Volumes III and IV of the Collected Papers of C. S. Peirce (1931–36) gives a number of logical diagrams or logical graphs which he says supports his Synechism. At the end of his logical discussion of Gamma Graphs (4.573 ff) Peirce says, “We here reach a point at which novel considerations about the constitution of knowledge and therefore of the constitution of nature burst in upon the mind with cataclysmical multitude and resistlessness. It is that synthesis of tychism and pragmatism for which I long ago proposed the name, Synechism, to which one thus returns; but this time with stronger reasons than ever before” (4.584). See also Existential Graphs, MS 514, by Charles Sanders Peirce with commentary by John F. Sowa at http://www.jfsowa.com/peirce/ms514

References

  • Aczel, A. D. (2003). Entanglement. New York: Plume.

    Google Scholar 

  • Adami, C. (2004). Information theory in molecular biology. Physics of Life Reviews, 1, 3–22.

    Article  Google Scholar 

  • Barbieri, M. (2003). The organic codes. Cambridge: The University Press.

    Google Scholar 

  • Barzun, J. (2000). From dawn to decadence. New York: HarperCollins.

    Google Scholar 

  • Bohr, N. (1935). Can quantum-mechanical description of physical reality be considered complete? Physics Review, 48, 696–702.

    Article  CAS  Google Scholar 

  • Clark, A. (1997). Being there. Putting brain, body, and world together again. Cambridge: MIT/Bradford.

    Google Scholar 

  • Delbrück, M. (1970). A physicist’s renewed look at biology: twenty years later. Science, 168, 1312–1315.

    Article  PubMed  Google Scholar 

  • Eddington, A. S. (1929). The nature of the physical world. Cambridge: The University Press.

    Google Scholar 

  • Eden, R. J. (1951). The quantum mechanics of non-holonomic systems. Proceedings of the Royal Society of London, 205A, 583–595.

    Google Scholar 

  • Eigen, M. (2002). Error catastrophe and antiviral strategy. PNAS, 99(21), 13374–13376.

    Article  PubMed  CAS  Google Scholar 

  • Emmeche, C. (1999). The biosemiotics of emergent properties in a pluralistic ontology. In E. Taborsky (Ed.), Semiosis, evolution, energy: Toward a reconceptualization of the sign (pp. 89–108). Achen: Shaker Verlag.

    Google Scholar 

  • Faísca, P. F. N. (2009). The nucleation mechanism of protein folding: A survey of computer simulation studies. J. Phys.: Condens. Matter, 21, 373102.

    Google Scholar 

  • Favareau, D. (2007). The evolutionary history of biosemiotics. In M. Barbieri (Ed.), Introduction to biosemiotics (pp. 1–68). Dordrecht: Springer.

    Chapter  Google Scholar 

  • Gell-Mann, M. (1994). The quark and the jaguar. New York: W. H. Freeman.

    Google Scholar 

  • Ghiselin, B. (1952). The creative process. Berkeley: The University of California Press. Paperback Edition 1985.

    Google Scholar 

  • Gilder, L. (2008). The age of entanglement. New York: Knopf.

    Google Scholar 

  • Godfrey-Smith, P. (2007). Information in biology. In D. Hull & M. Ruse (Eds.), The cambridge companion to the philosophy of biology (pp. 103–119). Cambridge: The University Press.

    Chapter  Google Scholar 

  • Goodwin, B. (1996). How the leopard changed its spots. New York: Touchstone.

    Google Scholar 

  • Hadamard. (1945). The psychology of invention in the mathematical field. Princeton: Princeton University Press.

    Google Scholar 

  • Hauser, M. D., Chomsky, N., & Fitch, W. T. (2002). The faculty of language: What is it, who has it, and how did it evolve? Science, 298, 1569–1579.

    Article  PubMed  CAS  Google Scholar 

  • Hertz, H. (1984). The principles of mechanics. Dover, NY, pp. 1–2. Original German ed., Prinzipien Mechanik, 1894.

  • Hoffmeyer, J. (2008). Biosemiotics. Scranton: University of Scranton Press.

    Google Scholar 

  • Hoffmeyer, J., & Emmeche, C. (1991). Code duality and the semiotics of nature. In M. Anderson & F. Merrell (Eds.), On semiotic modeling (pp. 117–166). NewYork: Mouton de Gruyter.

    Google Scholar 

  • Kauffman, S. A. (1993). The origins of order. Oxford: The University Press.

    Google Scholar 

  • Kelso, J. A. S. (1997). Dynamic patterns. Cambridge: MIT Press.

    Google Scholar 

  • Kull, K., Emmeche, C., & Favereau, D. (2008). Biosemiotic questions. Biosemiotics, 1(1), 41–55.

    Article  Google Scholar 

  • Landauer, R. & Bennett, C. H. (1985). The fundamental physical limits of computation. Scientific American, 48–56.

  • Leff, H. S., & Rex, A. F. (Eds.). (1990). Maxwell’s demon, entropy, information, computing. Princeton: Princeton University Press.

    Google Scholar 

  • Maynard-Smith, J. (1999). The Idea of Information in Biology. The Quarterly Review of Biology, 74(4), 395–400.

    Article  PubMed  CAS  Google Scholar 

  • Miller. (1984). Imagery in scientific thought. Boston: Birkhäuser.

    Google Scholar 

  • Orgel, L. E. (1963). The maintenance of the accuracy of protein synthesis and its relevance to aging. Proceedings of the National Academy of Science USA, 49, 517–21.

    Article  CAS  Google Scholar 

  • Pattee, H. H. (1967). Quantum mechanics. Heredity and the origin of life. Journal of Theoretical Biology, 17, 410–420.

    Article  PubMed  CAS  Google Scholar 

  • Zurek, W. H. (1990). Complexity, entropy, and the physics of information. Redwood City: Addison-Wesley.

    Google Scholar 

  • Pattee, H. H. (1968). The physical basis of coding and reliability in biological evolution. In Towards a theoretical biology, 1, prolegomena, C. H. Waddington (Ed.), Edinburgh University Press, pp. 67–93.

  • Pattee, H. H. (1969). How does a molecule become a message? Developmental Biology Supplement, 3, 1–16.

    Google Scholar 

  • Wigner, E. P. (1967). The problem of measurement. In Symmetries and Reflections. Bloomington: The University Press.

    Google Scholar 

  • Pattee, H. (1971). Can Life Explain Quantum mechanics? In T. Bastin (Ed.), Quantum theory and beyond — Essays and discussions arising from a colloquium (pp. 307–319). Cambridge: The University Press.

    Google Scholar 

  • Wheeler, J. A. (1983). Law without law. In J. A. Wheeler & W. H. Zurek (Eds.), Quantum theory and measurement (pp. 182–213). Princeton: Princeton University Press.

    Google Scholar 

  • Pattee, H. H. (1972). Physical problems of decision-making constraints. International Journal of Neuroscience, vol. 3, pp. 99–106. Also in Physical Principles of Neuronal and Organismic Behavior, M. Conrad and M. Magar, (Eds.), New York: Gordon and Breach, 1973, pp. 217–223.

  • Pattee, H. H. (1974). Discrete and continuous processes in computers and brains. In M. Conrad, W. Guttinger, & M. DalCin (Eds.), The physics and mathematics of the nervous system (pp. 128–148). New York: Springer.

    Chapter  Google Scholar 

  • von Weizsäcker, C. F. (1941). Zur deutung der Quantenmechanik. Zeitschrift für Physik, 118, 489–509.

    Article  Google Scholar 

  • Pattee, H. (1989). Clues from molecular symbol systems. In U. Bellugi & M. Studdert-Kennedy (Eds.), Signed and Spoken Language: Biological Constraints on Linguistic Form (pp. 261–274). Weinheim: Verlag Chemie GmbH. Dahlem Konferenzen 1980.

    Google Scholar 

  • Pattee, H. H. (1997). The Physics of Symbols and the Evolution of Semiotic Controls. Paper presented at the Workshop on Control Mechanisms for Complex Systems: Issues of Measurement and Semiotic Analysis, Las Cruces, New Mexico, December 8–12, 1996. (Note: My reference incorrectly states, “To be published as a Santa Fe Institute Study.” As far as I can tell, this paper was never published.)

  • Pattee, H. H. (2000). Causation, control, and the evolution of complexity. In Downward Causation, P. B. Anderson, C. Emmeche, N. O. Finnemann, and P. V. Christiansen (Eds.), Aarhus University Press, pp. 63–77.

  • Vedral, V. (2011). Living in a quantum world. Scientific American, 304(6), 38–43.

    Article  PubMed  Google Scholar 

  • Pattee, H. H. (2001). The physics of symbols: Bridging the epistemic cut. Bio Systems, 60, 5–21.

    Article  PubMed  CAS  Google Scholar 

  • Pauli, W. (1994). The philosophical significance of the idea of complementarity. In C. P. Enz & K. von Meyenn (Eds.), Writings on physics and philosophy (pp. 35–48). Berlin: Springer-Verlag.

    Chapter  Google Scholar 

  • Peirce, C. S. (1931–36). In C. Hartshorne & P. Weiss (Eds.), The collected papers. Volumes 1–6. Cambridge: Harvard University Press.

    Google Scholar 

  • Pinker, S., & Jackendoff, R. (2005). The nature of the language faculty and its implications for evolution of language (Reply to Fitch, Hauser, and Chomsky). Cognition, 97, 211–225.

    Article  Google Scholar 

  • Polanyi, M. (1968). Life’s irreducible structure. Science, 160, 1308–1312.

    Article  PubMed  CAS  Google Scholar 

  • Polanyi, M. & Prosch, H. (1975). Meaning. The University of Chicago Press.

  • Port, R., & van Gelder, T. (1995). Mind as motion: Explorations in the dynamics of cognition. Cambridge: MIT/Bradford.

    Google Scholar 

  • Quastler, H. (1953). Information theory in biology. Urbana: University of Illinois Press.

    Google Scholar 

  • Rocha, L. M. (2001). Evolution with material symbol systems. Bio Systems, 60, 95–121.

    Article  PubMed  CAS  Google Scholar 

  • Schneider, T. D. (2000). Evolution of Biological Information. Nucleic Acid Research 28(14), 2794–2799. http://www-lmmb.ncifcrf.gov/~toms/paper/ev/ev.pdf .

  • Schuster, P. (2006). Prediction of RNA secondary structures: From theory to models and real molecules. Reports on Progress in Physics, 69, 1419–1477.

    Article  CAS  Google Scholar 

  • Sereno, M. I. (1991). Four analogies between biological and cultural/linguistic evolution. Journal of Theoretical Biology, 151, 467–507.

    Article  PubMed  CAS  Google Scholar 

  • Simon, H. A. (1962). The architecture of complexity: Hierarchical systems. Proc. American Philosophical Soc. 106, 467–482. Reprinted in H. A. Simon, The Sciences of the Artificial, Chap. 8. Cambridge, MA: MIT Press (1996).

  • Sommerfeld, A. (1952). Lectures on theoretical physics, vol. 1, Mechanics (p. 50). New York: Academic.

    Google Scholar 

  • Sowa, J. F. (2000). Knowledge representation. Pacific Grove: Brooks/Cole.

    Google Scholar 

  • Stent, G. (1968). That was the molecular biology that was. Science, 160, 390–395.

    Article  PubMed  CAS  Google Scholar 

  • Szathmáry, E. (1992). What is the optimum size for the genetic alphabet? Proceedings of the National Academy of Sciences, USA, 89, 2614–2618.

    Article  Google Scholar 

  • von Neumann, J. (1955). Mathematical foundations of quantum mechanics (translated from the German by Robert T. Beyer) Princeton University Press, pp. 419–420.

  • von Neumann, J. (1966). In A. W. Burks (Ed.), Theory of self-reproducing automata (pp. 74–87). Urbana: University of Illinois Press. Lecture 5.

    Google Scholar 

  • Wheeler, J. A., & Zurek, W. H. (Eds.). (1983). Quantum theory and measurement. Princeton: Princeton University Press.

    Google Scholar 

  • Wolynes, P. G., Onuchic, J. N., & Thirumalai, D. (1995). Navigating the folding routes. Science, 267, 1619–1620.

    Article  PubMed  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to H. H. Pattee.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pattee, H.H. Epistemic, Evolutionary, and Physical Conditions for Biological Information. Biosemiotics 6, 9–31 (2013). https://doi.org/10.1007/s12304-012-9150-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12304-012-9150-8

Keywords

Navigation