Skip to main content

Constraints and Organisational Closure

  • Chapter
Biological Autonomy

Part of the book series: History, Philosophy and Theory of the Life Sciences ((HPTL,volume 12))

  • 1426 Accesses

Abstract

The aim of this first chapter is to propose a theoretical and formal framework that characterises closure as a causal regime specifically at work in biological organisation. In particular, it will be our contention that biological systems can be shown to involve two distinct, although closely interdependent, regimes of causation: an open regime of thermodynamic processes and reactions, and a closed regime of dependence between components working as constraints.

Some of the ideas exposed in this chapter, as well as some parts of the text, are taken from (Montévil and Mossio 2015).

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 89.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 119.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 119.99
Price excludes VAT (USA)
  • Durable hardcover 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

Notes

  1. 1.

    Strictly speaking, “mutual dependence” and “closure” are not synonymous. While the former is realised by any (sub)set of entities which depend on each other, the latter is realised by the set of all entities which are mutually dependent in a system. So for instance, the heart and the lungs realise mutual dependence among them, but only the whole set of organs of the organism realises (by hypothesis) closure.

  2. 2.

    Another, more complex, example is the atmospheric reaction networks, which realise a closed loop of chemical reactions (Centler and Dittrich 2007).

  3. 3.

    It should be noted that, over the years, Varela himself has proposed slightly different definitions of operational closure. Also, more recent contributions have introduced a theoretical distinction between organisational and operational closure: whereas “organisational” closure indicates the abstract network of relations that defines the system as a unity, “operational” closure refers to the recurrent dynamics and processes of such a system (see Thompson 2007).

  4. 4.

    According to Varela, three realisations of closure have been described: the cell, the immune system, and the nervous system (see Varela 1981: 18).

  5. 5.

    As a matter of fact, some authors have recently argued that the requirement for a physical boundary should be replaced by one for a functional boundary (Bourgine and Stewart 2004; Zaretzky and Letelier 2002). We agree entirely with this suggestion (see also Sect. 1.6 below), but it should be noted that functional boundaries, given that they are more general, might expose even more closure to the danger of applying to irrelevant systems. The appeal to functional boundaries should then go with a more rigorous definition of closure.

  6. 6.

    The connection between closure and constraints has been already put forward in the work of authors like Bickhard, Christensen, Hooker, and Kauffman, mentioned in the Introduction. Similarly, substantial theoretical work has been done on this issue by various members of the IAS Research Group over the last two decades.

  7. 7.

    A terminological clarification: For reasons that will become clearer later on (in particular in Chap. 5), we hold that biological self-determination (as self-constraint) implies specifically “self-maintenance” and not “self-generation”. Biological systems maintain themselves but do not generate themselves spontaneously (as wholes, although of course they do generate some their functional components). In this book, we will then use “self-maintenance” to refer to the specific mode of biological self-determination.

  8. 8.

    In our knowledge, (Piaget 1967) was the first author who has explicitly expressed the conceptual distinction between organisational closure and thermodynamic openness. The treatment of the distinction developed in this chapter is consistent, we think, with his own conception.

  9. 9.

    The terms implementation and realisation are often used to denote a very similar meaning. However, in a strict sense, an implementation is interpreted as a kind of physical realisation of a given formal organisation which is not unique (i.e., where there are multiple possibilities of realisation of said organisation: e.g., a computer programme can be completely specified in an abstract way and then implemented in various kinds of hardware). Consequently, in this book, when we want to avoid such an interpretation, we shall use the term (physical) realisation.

  10. 10.

    In a cohesive or coherent movement of constituents in the system, for instance, the classic idea of mechanical work.

  11. 11.

    Typically, in a molecular bond, related to chemical work.

  12. 12.

    ‘Heat’ refers to energy that is disordered relative to the initial state of the current exothermic transition. Only in that situation does the fact of “not being recoverable any more” have a clear meaning. As we see, the concepts of work and heat are defined in terms of possibilities of energy use. But “use” here refers to a functionality which is clearly external to the system; hence, it lacks all significance without the presence of an outside observer. Insofar as living organisms are autonomous systems, we shall have to restate these concepts in such a way that they acquire meaning within the operational framework of the actual system (in Chap. 3, we shall discuss this question in detail).

  13. 13.

    Actually, (Atkins 1984) does not speak about self-maintenance or biology. Rather, his book is about a general interpretation of thermodynamics, and the asymmetry between heat and work (work as a “constrained release of energy”).

  14. 14.

    On the one hand, the system must couple with some external source of energy (sunlight or chemical energy in the autotrophic case; extraneous organic matter in the heterotrophic one). On the other hand, it is also fundamental that internal energetic couplings take place, because this allows certain processes (of synthesis, typically) to occur at the expense of others (degradation), when in principle the former ones alone would not be spontaneously viable.

  15. 15.

    Feasible in the sense that, when coupled, a global decrease of free energy should take place.

  16. 16.

    Originally, a work cycle is a set of externally controlled processes that takes a thermodynamic system back to its initial state, giving as a result an overall production or consumption of work. The typical thermodynamic system would be a gas enclosed in a thermal machine (with walls that can be adiabatic or kept at constant temperature if required) undergoing successive expansions or compressions until it is brought back to its original thermodynamic state. The Carnot cycle in particular is completed through two isothermal and two adiabatic processes, producing ideally an amount of work that equals (or is exactly proportional to) the area limited by the lines representing those processes in a pressure-volume diagram.

  17. 17.

    The latter precision is important because it would otherwise be trivially true that a situation AB and a situation ACB are different, because of the new object (C) that has been added. Yet, the presence of C does not necessarily change something for the objects present only in the first situation (A and B), since this depends on whether they interact with C in a relevant way.

  18. 18.

    It is crucial to stress that the conservation concerns only these relevant aspects, while other aspects of the entity that exerts the constraint might undergo alteration, even at τ.

  19. 19.

    The definition of constraint provided above is reminiscent of (and, we think, consistent with) Pattee’s account of this concept (see for example, Pattee 1972, 1973). This author defines a constraint as an “alternative description” of the dynamical behaviour of a system, in which a macroscopic material structure selectively limits the degrees of freedom of a local microscopic system. For an extensive discussion of Pattee’s account, see also (Umerez 1994, 1995).

  20. 20.

    Note that the conservation supposes that a specific time scale τ, at which the target process occurs, is to be specified which, in turn, requires determining when the process begins and ends. As a consequence, in those cases in which the process is continuously occurring, discretisation might be necessary to describe the constraints. Let’s take the physical example of a river continuously eroding its banks. At first sight, the banks could not be taken, according to our definition, as constraints on the dynamics of the river, precisely because they are transformed by the river. But in fact this description of the system is inadequate, because it fails in specifying the relevant time scale. Although the banks are of course not conserved at the very long time scale at which the entire existence of the river can be described, their relevant aspects by virtue of which the river (i.e. a specifiable set of water molecules) moves from a specific point upstream to a specific point downstream in given period of time are presumably conserved during that period. Accordingly, the banks, at that time scale, fit our definition.

  21. 21.

    At biologically relevant time scales, then, the distinction between constraints and processes roughly maps onto Rosen’s distinction between efficient and material causes (Rosen 1991): constraints might indeed be said to “efficiently” produce an effect by acting, for instance, on the underlying “material” input of a reaction. In spite of this (approximate) correspondence, however, we do not adopt Rosen’s terminology, which can be confusing in some respect (see also Pattee 2007 on this point), and will maintain in this book the distinction between constraints and processes. Actually, it might be argued that constraints should rather be intended as “formal” causes (see for example Emmeche et al. 2000; we also briefly discuss this question in Chap. 2).

  22. 22.

    As we will discuss at length in Chap. 5, due to the degree of complexity required by autonomous systems, these can only be historical systems (i.e. systems whose complexity has emerged through a cumulative phylogenetic process), and can by no means appear spontaneously (as dissipative structures do).

  23. 23.

    In the case of repair the entity is maintained, while in the case of replacement it is destroyed and reconstructed. Note that the same situation can be interpreted as a case of replacement or repair following the scale at which the constraint is described: individual enzymes are replaced, while the population is repaired. This holds for all those cases (mainly at the molecular level) in which both individual and populations exert the same constraint. See the discussion about scale invariant constraint in Sect. 1.6 below.

  24. 24.

    See also note 1 above on the conceptual relations between “closure” and “mutual dependence”.

  25. 25.

    The relations brought about by constraints responsible for closure in living systems have received two characterisations by Howard Pattee, in different stages of his work: statistical closure (1973) and semantic closure (Pattee 1982). By “statistical closure” he (1973: 94–97) means a collection of elements that may combine or interact with each other individually in many ways, but that nevertheless persists as the same collection largely because of the rates of their combination. This in turn implies a population dynamics for the elements and therefore a real-time dependence. Furthermore, the rates of specific combinations of elements must be controlled by collections of the elements of the closed set. The adjective statistical refers to the “selective loss of detail” of a statistical classification presents in relation to the underlying dynamics. It explains, according to Pattee, the nature and function of control constraints within a hierarchical system.

    In turn, Pattee defines “semantic closure” as follows: “We can say that the molecular strings of the genes only become symbolic representations if the physical symbol tokens are, at some stage of string processing, directly recognized by translation molecules (tRNA’s and synthetases) which thereupon execute specific but arbitrary functions (protein synthesis). The semantic closure arises from the necessity that the translation molecules are themselves referents of gene strings.” (Pattee 1982: 333). Semantic closure is then based on the idea of symbolic records that preserve those constraints, and of how they are interpreted within the living system as a whole (Umerez 1995; Etxeberria and Moreno 2001).

  26. 26.

    Scale-invariant constraints may be realised in the form of both redundancy or degeneracy of functional parts. As (Tononi et al. 1999) have pointed out, redundancy refers to the situation in which structurally similar elements produce the same effects, whereas degeneracy occurs when structurally different elements perform the same function.

  27. 27.

    It is, of course, conceivable that a description of constraints might possibly be given in terms of thermodynamics, specifically as entities that are not affected by the thermodynamic flow. However, in this case, constraints (and hence closure) would not be reduced to a different causal regime, but simply re-described in different terms.

  28. 28.

    This implication allows us to distinguish between a closure of constraints and a cycle of processes or reactions such as, for instance, the hydrologic cycle mentioned in the introduction to this chapter. In this case, the entities involved (e.g. clouds, rain, springs, rivers, seas, clouds, etc.) are connected to each other in such a way that they generate a cycle of transformations and changes between them. In turn, these entities do not act as constraints on each other (among other reasons, precisely because they are transformed when they produce another water structure), and the system can be adequately described by appealing to a set of external boundary conditions (soil, sun, etc.) acting on a single causal regime of thermodynamic changes (see also Mossio and Moreno 2010).

  29. 29.

    See also (Montévil and Mossio 2015) for more details in this issue. In Chap. 3, Sect. 3.3.1 below we will explore the issue of the temporal boundaries of closure, when these go beyond the lifespan of an individual organism.

  30. 30.

    We made a contribution in Mossio et al. (2009a), in which we analysed one of Rosen’s claims, according to which closure to efficient causation has non-computable models. (Cárdenas et al. 2010) offers a detailed reply to our analysis.

  31. 31.

    It should be mentioned that there is another conception of minimal metabolism, typically put forward by biochemists (see for instance Gil et al. 2004), according to which it is the characterisation of “minimal genomes” through the simplification of existing ones, under the assumption that their associated metabolic networks will drastically reduce the complexity of extant metabolisms. Here, minimal metabolisms are still “genetically-instructed metabolisms”, similar (although highly simplified) to those realised by fully-fledged living organisms. As discussed by (Morowitz 1992) and (Morange 2003), one of the problems of this conception seems to be that, since metabolic simplicity depends on the environment, it is highly problematic to elaborate shared criteria to determine what the minimal metabolic network actually is.

  32. 32.

    For example, (Eschenmosser 2007: 311), writes that “another type of reaction loop that can emerge as a consequence of the exploration of a chemical environment’s structure and reactivity space is one that, driven by the free energy of starting materials, connects intermediate products (substrates as opposed to catalysts) in a cyclic pathway: such a cycle is referred to as autocatalytic metabolic cycle.”

  33. 33.

    (De Duve 2007) uses the term “proto-metabolism” to denote those chemical networks driven by catalysts that, whatever their nature, cannot have displayed the exquisite specificity of present-day enzymes and must necessarily have produced some sort of “dirty gemisch”.

  34. 34.

    We focus here on deleterious variations, i.e. variations that do not lead to new viable organisations and would disrupt the system if not compensated.

  35. 35.

    It should be noted that, in some cases, variations may be neutral with regard to the self-maintenance of the system: in spite of the variation, the system may drift, but closure is conserved. And it might be the case that the biological system exerts a form of compensation even on this kind of harmless variation, counteracting its effects. In what follows, however, we shall not discuss these forms of compensation because they are negligible with respect to maintaining closure, which is, after all, the main reason for requiring regulation. Regulation will then specifically be characterised in relation to cases of “deleterious” variations that disrupt closure: a gap is generated between the conditions of existence and the activity of the system, which is no longer able to meet those very conditions of existence, and is therefore destined to collapse.

  36. 36.

    The content of this section owes a lot to preliminary discussions with Leonardo Bich and Kepa Ruiz-Mirazo. See Bich et al. (forthcoming) for details.

  37. 37.

    As Di Paolo puts it, adaptivity is “a system’s capacity to regulate, according to the circumstances, its states and its relation to the environment with the result that, if the states are sufficiently close to the boundary of viability, (1) tendencies are distinguished and acted upon depending on whether the states will approach or recede from the boundary and, as a consequence, (2) tendencies of the first kind are moved closer to or transformed into tendencies of the second and so future states are prevented from reaching the boundary with an outward velocity” (Di Paolo 2005: 438). For a detailed discussion, see also (Barandiaran et al. 2009); (Barandiaran and Egbert 2013).

  38. 38.

    (Bechtel 2007) has pointed out the same argument, which he states as follows: “if control is to involve more than strict linkage between components, what is required is a property in the system that varies independently of the basic operations. The manipulation of this property by one component can then be coordinated with a response to it by another component so that one component can exert control over the operation of the other component ” (Bechtel 2007: 290).

  39. 39.

    It is worth emphasising that, from the autonomous perspective, biological organisations might be hierarchical with respect to both orders and levels of closure. Chapter. 6 addresses explicitly this conceptual distinction.

References

  • Atkins, P. W. (1984). The second law. New York: Freeman.

    Google Scholar 

  • Barandiaran, X., & Egbert, M. D. (2013). Norm-establishing and norm-following in autonomous agency. Artificial Life Journal, 20(1), 5–28.

    Article  Google Scholar 

  • Barandiaran, X., Di Paolo, E., & Rohde, M. (2009). Defining agency. Individuality, normativity, asymmetry and spatio-temporality in action. Journal of Adaptive Behavior, 17(5), 367–386.

    Article  Google Scholar 

  • Bechtel, W. (2007). Biological mechanisms: organized to maintain autonomy. In F. Boogerd, F. Bruggeman, J. H. Hofmeyr, & H. V. Westerhoff (Eds.), Systems biology. Philosophical foundations (pp. 269–302). Amsterdam: Elsevier.

    Chapter  Google Scholar 

  • Bich, L., Moreno, M., Mossio, M., & Ruiz-Mirazo, K. (forthcoming). Biological regulation.

    Google Scholar 

  • Bickhard, M. H. (2000). Autonomy, function, and representation. Communication and Cognition Artificial Intelligence, 17(3–4), 111–131.

    Google Scholar 

  • Bourgine, P., & Stewart, J. (2004). Autopoiesis and cognition. Artificial Life, 10, 327–345.

    Article  Google Scholar 

  • Bourgine, P., & Varela, F. J. (1992). Toward a practice of autonomous systems. Cambridge, MA: MIT Press/Bradford Books.

    Google Scholar 

  • Cárdenas, M. L., Letelier, J. C., Gutierrez, C., Cornish-Bowden, A., & Soto-Andrade, J. (2010). Closure to efficient causation, computability and artificial life. Journal of Theoretical Biology, 263(1), 79–92.

    Article  Google Scholar 

  • Centler, F., & Dittrich, P. (2007). Chemical organizations in atmospheric photochemistries: a new method to analyze chemical reaction networks. Planetary and Space Science, 55(4), 413–428.

    Article  Google Scholar 

  • Chandler, J. L. R., & Van De Vijver, G. (Eds.). (2000). Closure: emergent organizations and their dynamics (Vol. 901). New York: Annals of the New York Academy of Science.

    Google Scholar 

  • Chandresekhar, S. (1961). Hydrodynamic and hydromagnetic stability. Oxford: Clarendon.

    Google Scholar 

  • Christensen, W. D. (2012). Natural sources of normativity. Studies in History and Philosophy of Science Part C, 43(1), 104–112.

    Article  Google Scholar 

  • Christensen, W. D., & Hooker, C. A. (2000). An interactivist-constructivist approach to intelligence: self-directed anticipative learning. Philosophical Psychology, 13, 5–45.

    Article  Google Scholar 

  • De Duve, C. (2007). Chemistry and selection. Chemistry and Biodiversity, 4(4), 574–583.

    Article  Google Scholar 

  • Deamer, D. W. (2009). On the origin of systems. Systems biology, synthetic biology and the origin of life. EMBO Reports, 10(1), S1–S4.

    Article  Google Scholar 

  • Di Paolo, E. A. (2005). Autopoiesis, adaptivity, teleology, agency. Phenomenology and the Cognitive Sciences, 4(4), 429–452.

    Article  Google Scholar 

  • Emmeche, C. (1992). Life as an abstract phenomenon: is artificial life possible? In F. Varela & P. Bourgine (Eds.), Toward a practice of autonomous systems (pp. 466–474). Cambridge, MA: MIT Press.

    Google Scholar 

  • Emmeche, C., Stjernfelt, F., & Køppe, S. (2000). Levels, emergence, and three versions of downward causation. In P. B. Andersen, C. Emmeche, N. O. Finnemann, & P. V. Christiansen (Eds.), Downward causation. minds, bodies and matter (pp. 13–34). Aarhus: Aarhus University Press.

    Google Scholar 

  • Eschenmosser, A. (2007). The search for the chemistry of life’s origin. Tetrahedron, 63, 12821–12844.

    Article  Google Scholar 

  • Etxeberria, A., & Moreno, A. (2001). From complexity to simplicity: nature and symbols. BioSystems, 60(1–3), 149–157.

    Article  Google Scholar 

  • Farmer, J., Kauffman, S., & Packard, N. (1986). Autocatalytic replication of polymers. Physica D, 22, 50–67.

    Article  Google Scholar 

  • Fontana, W. (1992). Algorithmic chemistry. In C. G. Langton, C. Taylor, J. D. Farmer, & S. Rasmussen (Eds.), Artificial life II (pp. 159–209). Redwood City: Addison-Wesley.

    Google Scholar 

  • Fontana, W., Wagner, G., & Buss, L. W. (1994). Beyond digital naturalism. Artificial Life, 1(1/2), 211–227.

    Google Scholar 

  • Ganti, T. (1975). Organization of chemical reactions into dividing and metabolizing units: the chemotons. BioSystems, 7, 15–21.

    Article  Google Scholar 

  • Gil, R., Silva, F. J., Peretó, J., & Moya, A. (2004). Determination of the core of a minimal bacteria gene set. Microbiology and Molecular Biology Reviews, 68, 518–537.

    Article  Google Scholar 

  • Glansdorff, P., & Prigogine, I. (1971). Thermodynamics of structure, stability and fluctuations. London: Wiley.

    Google Scholar 

  • Griesemer, J. R., & Szathmáry, E. (2009). Ganti’s chemoton model and life criteria. In S. Rasmussen, M. A. Bedau, L. Chen, D. Deamer, D. C. Krakauer, & N. H. Packard (Eds.), Protocells: bridging nonliving and living matter (pp. 481–512). Cambridge, MA: MIT Press.

    Google Scholar 

  • Juarrero, A. (1999). Dynamics in action: intentional behavior as a complex system. Cambridge, MA: MIT Press.

    Google Scholar 

  • Kauffman, S. (1986). Autocatalytic sets of proteins. Journal of Theoretical Biology, 119, 1–24.

    Article  Google Scholar 

  • Kauffman, S. (2000). Investigations. Oxford: Oxford University Press.

    Google Scholar 

  • Letelier, J.-C., Marin, J., & Mpodozis, J. (2003). Autopoietic and (M, R)-systems. Journal of Theoretical Biology, 222, 261–272.

    Article  Google Scholar 

  • Letelier, J.-C., Soto-Andrade, J., Guiñez Abarzua, F., Cornish-Bowden, A., & Cardenas, M. L. (2006). Organizational invariance and metabolic closure: analysis in terms of (M, R)-systems. Journal of Theoretical Biology, 238, 949–961.

    Article  Google Scholar 

  • Longo, G., & Montevil, M. (2014). Perspectives on organism: biological time, symmetries and singularities. Heidelberg: Springer.

    Book  Google Scholar 

  • Mattick, J. (2004). The hidden genetic program of complex organisms. Scientific American, 291(4), 60–67.

    Article  Google Scholar 

  • Maynard Smith, J. (1986). The problems of biology. Oxford: Oxford University Press.

    Google Scholar 

  • McMullin, B. (1997). SCL: an artificial chemistry in Swarm. Santa Fe Institute Working Paper (SFI).

    Google Scholar 

  • McMullin, B., & Varela, F. J. (1997). Rediscovering computational autopoiesis. In P. Husbands & I. Harvey (Eds.), Proceedings of the fourth European conference on artificial life (pp. 38–47). Cambridge, MA: MIT Press.

    Google Scholar 

  • Montévil, M., & Mossio, M. (forthcoming). Closure of constraints in biological organisation.

    Google Scholar 

  • Morán, F., Moreno, A., Montero, F., & Minch, E. (1997). Further steps towards a realistic description of the essence of life. In C. G. Langton & K. Shimohara (Eds.), Artificial life V (Proceedings of the 5th international workshop on the synthesis and simulation of living systems) (pp. 255–263). London: MIT Press.

    Google Scholar 

  • Morange, M. (2003). La vie expliquée ? 50 ans après la double hélice. Paris: Odile Jacob. Eng trans. (2008). Life explained. New Haven: Yale University Press.

    Google Scholar 

  • Moreno, A., & Ruiz Mirazo, K. (1999). Metabolism and the problem of its universalization. BioSystems, 49(1), 45–61.

    Article  Google Scholar 

  • Moreno, A., Etxeberria, A., & Umerez, J. (1994). Universality without matter? In R. A. Brooks & P. Maes (Eds.), Artificial life IV (Proceedings of the 4th international workshop on the synthesis and simulation of living systems) (pp. 406–410). London: MIT Press.

    Google Scholar 

  • Morowitz, H. J. (1968). Energy flow in biology. New York: Academic Press.

    Google Scholar 

  • Morowitz, H. J. (1992). Beginnings of cellular life. New Haven: Yale University Press.

    Google Scholar 

  • Mossio, M. (2013). Closure, causal. In W. Dubitzky, O. Wolkenhauer, K.-H. Cho, & H. Yokota (Eds.), Encyclopedia of systems biology. New York: Springer.

    Google Scholar 

  • Mossio, M., & Moreno, A. (2010). Organizational closure in biological organisms. History and Philosophy of the Life Sciences, 32(2–3), 269–288.

    Google Scholar 

  • Nicolis, G., & Prigogine, I. (1977). Self-organisation in non-equilibrium systems: from dissipative structures to order through fluctuation. New York: Wiley.

    Google Scholar 

  • Pattee, H. H. (1972). Laws and constraints, symbols and languages. In C. H. Waddington (Ed.), Towards a theoretical biology (Vol. 4). Edinburgh: Edinburgh University Press.

    Google Scholar 

  • Pattee, H. H. (1977). Dynamic and linguistic modes of complex systems. International Journal of General Systems, 3, 259–266.

    Article  Google Scholar 

  • Pattee, H. H. (1982). Cell psychology: an evolutionary approach to the symbol-matter problem. Cognition and Brain Theory, 4, 325–341.

    Google Scholar 

  • Pattee, H. H. (2007). Laws, constraints, and the modeling relation – History and interpretations. Chemistry and Biodiversity, 4, 2272–2295.

    Article  Google Scholar 

  • Piaget, J. (1967). Biologie et connaissance. Paris: Éditions de la Pléiade.

    Google Scholar 

  • Piedrafita, G., Montero, F., Morán, F., Cárdenas, M. L., & Cornish-Bowden, A. (2010). A simple self-maintaining metabolic system: robustness, autocatalysis, bistability. PLoS Computational Biology, 6(8), e1000872.

    Article  Google Scholar 

  • Rosen, R. (1991). Life itself. A comprehensive enquiry into the nature, origin and fabrication of life. New York: Columbia University Press.

    Google Scholar 

  • Ruiz-Mirazo, K. (2001). Condiciones físicas para la aparición de sistemas autónomos con capacidades evolutivas abiertas. Ph.D. dissertation, University of the Basque Country.

    Google Scholar 

  • Ruiz-Mirazo, K., & Mavelli, F. (2007). Simulation model for functionalized vesicles: lipid-peptide integration in minimal protocells. In F. Almeida e Costa, L. M. Rocha, I. Harvey, & A. Coutinho (Eds.), ECAL 2007, Lisbon, Portugal, 10–14 Sept 2007. Proceedings (Lecture notes in computer science 4648, pp. 32–41). Heidelberg: Springer.

    Google Scholar 

  • Ruiz-Mirazo, K., & Mavelli, F. (2008). Towards ‘basic autonomy’: stochastic simulations of minimal lipid- peptide cells. BioSystems, 91(2), 374–387.

    Article  Google Scholar 

  • Ruiz-Mirazo, K., & Moreno, A. (2004). Basic autonomy as a fundamental step in the synthesis of life. Artificial Life, 10(3), 235–259.

    Article  Google Scholar 

  • Szathmary, E. (2006). The origin of replicators and reproducers. Philosophical Transactions of the Royal Society B., 361, 1761–1776.

    Article  Google Scholar 

  • Thompson, E. (2007). Mind in life. Biology, phenomenology, and the sciences of mind. Cambridge, MA: Harvard University Press.

    Google Scholar 

  • Tononi, G., Sporns, O., & Edelman, G. M. (1999). Measures of degeneracy and redundancy in biological networks. Proceedings of the National Academy of Sciences USA, 96, 3257–3262.

    Google Scholar 

  • Umerez, J. (1995). Semantic closure: a guiding notion to ground artificial life. Advances in artificial life, 929, 77–94.

    Article  Google Scholar 

  • Umerez, J. (2001). Howard Pattee’s theoretical biology—a radical epistemological stance to approach life, evolution and complexity. Biosystems, 60(1), 159–177.

    Article  Google Scholar 

  • Van Dujin, M., Keijzer, F., & Franjen, D. (2006). Principles of minimal cognition: casting cognition as sensorimotor coordination. Adaptive Behavior, 14(2), 157–170.

    Article  Google Scholar 

  • Varela, F. J. (1981). Autonomy and autopoiesis. In G. Roth & H. Schwegler (Eds.), Self-organizing systems: an interdisciplinary approach (pp. 14–24). Frankfurt/New York: Campus Verlag.

    Google Scholar 

  • Varela, F. J. (1997). Patterns of life: intertwining identity and cognition. Brain and Cognition, 34, 72–87.

    Article  Google Scholar 

  • Vasas, V., Fernando, C., Santos, M., & Kauffman, S. (2012). Evolution before genes. Biol Direct, 7–1.

    Google Scholar 

  • Zaretzky, A., & Letelier, J. C. (2002). Metabolic networks from (M,R) systems and autopoiesis perspective. Journal of Biological Structures, 10(3), 265–280.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer Science+Business Media Dordrecht

About this chapter

Cite this chapter

Moreno, A., Mossio, M. (2015). Constraints and Organisational Closure. In: Biological Autonomy. History, Philosophy and Theory of the Life Sciences, vol 12. Springer, Dordrecht. https://doi.org/10.1007/978-94-017-9837-2_1

Download citation

Publish with us

Policies and ethics