Skip to main content
Log in

Integrated-structure emergence and its mechanistic explanation

  • Non-Standard Approaches to Emergence
  • Published:
Synthese Aims and scope Submit manuscript

Abstract

This paper proposes an integrated-structure notion of interlevel (part-whole) emergence, from a dynamic relational ontological perspective. First, I will argue that only the individualist essentialism of atomistic metaphysics can block the possibility of interlevel emergence. Then I will show that we can make sense of emergence by recognizing the formation of structures of transformative and interdependent causal relations in the generation and development of a particular class of mereological complexes called integrated systems. Finally, I shall argue that even though the emergent structural attributes of such systems are not micro-determined nor micro-reducible, they can still be accounted for by an interlevel integrative neo-mechanistic form of explanation.

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.

Fig. 1

Similar content being viewed by others

Notes

  1. See, for example, Mark Bedau and Paul Humphrey’s characterizations of emergence: “(1) Emergent phenomena are somehow constituted by, and generated from, underlying processes; (2) Emergent phenomena are somehow autonomous from underlying processes” (Bedau 1997: p. 376); “Emergence is, broadly speaking, the view that there are features of the world—objects: properties, laws: perhaps other things—that are manifested as a result of the existence of other, usually more basic, entities but that cannot be completely reduced to those other entities” (Humphreys 2006: p. 90).

  2. Following Auyang, I take here the notion of ‘quantitative’ as having “a more general meaning than that regarding amount or size. (…) An individual has various characters, the summary of which is called the individual’s state. The general concept of a character can be analyzed into two elements, a type and a value. Blue and red are values of the character type color; big and bigger, values of size. Other characters are temperature, weight, and annual income; their values − 25 °C, 250 kg, and $25,000—are the definite predicates attributed to individuals, and their types are represented in the units: degrees Celsius, kilogram, and dollar” (Auyang 1999: p. 177, 49).

  3. We owe to David Bohm one of the most accurate descriptions of the atomistic commitment, when he characterized the classical mechanistic view within Physics: “At bottom, the only changes that are regarded as possible within this scheme are quantitative changes in the parameters or functions (…), while fundamental qualitative changes in the modes of being of the basic entities and in the forms in which the basic laws are to be expressed are not regarded as possible. Thus, the essence of the mechanistic position lies in its assumption of fixed basic qualities, which means that the laws themselves will finally reduce to purely quantitative relationships” (Bohm 1984: p. 47, 131).

  4. Bitbol (2012) points out the importance of this distinction in the understanding of the behaviour and properties of a carbon atom given by quantum chemistry. Some of the chemical bonds the carbon atom can establish can be derived from the structure of the atom and can thus be accounted for by micro-reductionism. Still, in most cases, that is not possible. As Bitbol says: “If one starts from a quantum model of an isolated atom of carbon, one can predict neither its valence nor the geometry of its possible chemical bonds” (2012: p. 250). In other words, “the physical model of the isolated carbon atom taken as the starting point of bottom-up synthesis does not coincide with the chemical model of carbon-components of molecules that is the end product of top-down analysis. The bottom-up elements (2s2, \(2_{x}^{1} ,2_{y}^{1}\)-isolated carbons) do not coincide with the top-down components (sp3-hybridized carbons)” (Ibid.). Specifically, we need to know the other atoms with which that carbon atom will interact, the chemical bonds which will be established, the geometry of each specific molecule, and the relative position of the carbon atom within the molecular chain. I thank the anonymous reviewer who called this issue and Bitbol’s paper to my attention.

  5. For the fundamental difference between this notion of ‘internal relation’, originally elaborated by monist philosophers of the late nineteenth century, and the particular ways in which that notion was reformulated by Russell, Armstrong and Lewis, see Humphreys (2016: pp. 103–116), and Dunn (1990a, b).

  6. In this sense, my relational-transformative account of emergence differs from the transformational account that has been proposed by Humphreys (2016), Guay and Sartenaer (2016) and Sartenaer (2018), given their essentially ‘flat’ (‘same-level’ or ‘inter-domain’) perspective. Furthermore, such flat accounts do not consider the occurrence of relations or interactions as a necessary generating source of transformational processes of emergence. More recently, Anjum and Mumford (2017) also advocated a causal-transformative model of emergence, which is generally in line with the view proposed in Santos 2015a, albeit presented in the framework of their specific version of powers ontology and theory of causation.

  7. Against the ‘Simple view of aggregation’, Gillett endorses “the scientific emergentist’s ‘Conditioned view of aggregation’, which implies components have some powers in complex collectives that they would not have if the laws or principles applying in simpler collectives exhaustively applied in the complex aggregation” (2016: pp. 17–18).

  8. I borrow this notion of ‘structural conditioning’ from Archer (1995).

  9. The notion of mechanism can be defined as “a structure performing a function in virtue of its component parts, component operations, and their organization. The orchestrated functioning of the mechanism, manifested in patterns of change over time in properties of its parts and operations, is responsible for one or more phenomena” (Bechtel and Abrahamsen 2010: p. 323). Alternatively, Glennan and Illari proposed the following minimal notion of mechanism: “A mechanism for a phenomenon consists of entities (or parts) whose activities and interactions are organized so as to be responsible for the phenomenon” (2018: p. 2).

  10. According to Levins, an engineer’s circuit is a system in which the way the different parts (viz., units, condensers, transistors, wires, switches, etc.) are interrelated determines the system’s properties. Therefore, the properties of the system will not be derivable from simple statistics of the properties of individual parts, as it typically happens in the case of aggregate systems. In this kind of systems, the individual parts, although physically interacting with each other, “affect the properties of the whole only by virtue of being part of a mean, or a variance, as a part of frequency” (Levins 2011: p. 75). On the contrary, in simply composed or component systems, structure or organization matters. Still, “the properties of the parts do not affect the mode of response of each other, but only the way in which a signal is processed that passes through all of them” (Levins 2011: p. 76).

  11. Given the notion of integrated system, it is clear that the more the parts’ activities are mutually interconnected and interdependent, the more intertwined or entangled they become, and the less “successful is a sequential account of the mechanism in which each operation is treated as independent of the others” (Bechtel and Richardson 2010: p. xxxv). Therefore, it is natural to find some systems which seem to defy the classical mechanistic strategies of decomposition and localization. In such cases, the explanation of the system can only be referred to the global organization of its parts (2010: p. 27, 32), since parts “do not seem individually to contribute to anything of interest to understanding the behaviour of the whole” (2010: pp. 202–203). This fact raises, certainly, the problem of the limits of any mechanistic account, since the behaviour of those systems only seems to be accountable for by means of dynamical mathematical explanations, which do not refer to the micro-level interactions among the system’s parts, but rather characterize the development of the system in terms of its own higher-level variables. Silberstein and Chemero (2013) argued that, in systems neuroscience, explanations are not grounded on decomposition and localization, but rather on mathematical explanations in a more lawlike fashion. The question is then whether such alternative forms of explanation conflict with each other or can rather be seen as complementary. Bechtel, for example, has been defending the need of a new dynamic mechanistic explanation. I cannot, of course, pretend to have a conclusive answer to such a question. Still, I tend to favour the second alternative since the critical issue seems to me dependent on possible scientific and technological future developments (Bechtel and Richardson 2010: pp. xxxv–xxxvi; Kaplan 2018). The limits of the mechanistic approach must be thus taken as an open empirical question. We cannot settle the question a priori (but see: Silberstein 2020).

  12. This is why there is no substantial difference between the one–one (e.g., Wilson 2015; Baysan and Wilson 2017) and the one-many (Gillett 2016) readings of emergence relation. In my view, an emergent systemic attribute is both emergent from the collection or plurality of the parts in terms of their intrinsic properties and associated relations, as well as from the combinatorial-structural property which applies to that set.

  13. As Lewontin has so often emphasized, “DNA is a dead molecule, among the most nonreactive, chemically inert molecules in the living world. (…) DNA has no power to reproduce itself. Rather it is produced out of elementary materials by a complex cellular machinery of proteins. Although it is often said that DNA produces proteins: proteins (enzymes) produce DNA (…) Not only is DNA incapable of making copies of itself, aided or unaided, but it is incapable of ‘making’ anything else” (Lewontin and Levins 2007: p. 239). See also: Lewontin (2000) and Shapiro (2009).

  14. The debate about the precise relations between the notions of ‘constrain’ and ‘control’, as well as about their relations with the notion itself of ‘causation’, although highly relevant, goes well beyond the limits and the immediate goals of this paper. Still, see, for example: Juarrero (1999: pp. 131–162); Moreno and Mossio (2015: pp. 1–38); Winning and Bechtel (2018).

  15. From this perspective, I diverge from Gillett (2016), since his notion of downward causation (called ‘machresis’) is taken to be both non-causal and non-compositional.

  16. In this sense, I diverge from the transformational account proposed by Ganeri (2011), since in Ganeri’s view, emergent properties are instantiated by the transformed parts in their blending state, not by the system: “the elements themselves acquire new causal powers when they are in a certain state, namely the state of jointly composing a body: powers that they did not have beforehand when they were in other combinations with other elements. This is different from the view that the body as a whole has powers which none of its parts have individually. It is instead the view that the parts themselves have new powers conditionally upon their membership of the whole” (2011: p. 686). As Peter Simons rightly notes, it is important “to distinguish between a collection of many individuals and the one individual they compose, if they do” (2006: p. 599, n.4). From this perspective, my account fits the ‘new object’ type of response to the collapse problem as proposed by Baysan and Wilson (2017).

  17. As Bechtel lucidly observed, “mechanistic explanations are inherently reductionistic insofar as they require specifying the parts of a mechanism and the operations the parts perform. But they also require consideration of the organization of the whole mechanism and its relation to conditions in its environment since it is only when appropriately situated that a mechanism will produce the phenomenon of interest” (Bechtel 2011: p. 538—italics inserted). Therefore, “[t]he notion of reduction that arises with mechanistic explanation (…) is very different from that which has figured either in popular discussions or in recent philosophy of science, and its consequences are quite different. (…) While the functioning of a mechanism depends upon its constitution, it also depends on its context, including its incorporation within systems at yet higher levels of organization” (Bechtel 2006: pp. 40–41).

References

  • Alexander, S. (1920). Space, time, and deity. London: Macmillan.

    Google Scholar 

  • Anjum, R., & Mumford, S. (2017). Emergence and demergence. In M. P. Paoletti & F. Orilia (Eds.), Philosophical and scientific perspectives on downward causation (pp. 92–109). New York: Routledge.

    Chapter  Google Scholar 

  • Archer, M. (1995). Realist social theory: The morphogenetic approach. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Armstrong, D. (1978). Universals and scientific realism. Vol. II: A theory of universals. Cambridge: Cambridge University Press.

    Google Scholar 

  • Armstrong, D. (1997). A world of states of affairs. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Atlan, H. (2011). The genetic program. In H. Atlan (Ed.), Selected writings. On self-organization, philosophy, bioethics, and judaism (pp. 145–171). New York: Fordham University Press.

    Google Scholar 

  • Atlan, H., & Koppel, M. (1990). The cellular computer DNA: Program or data? Bulletin of Mathematical Biology, 52(3), 335–348.

    Article  Google Scholar 

  • Auyang, S. (1999). Foundations of complex-system theories – In economics, evolutionary biology, and statistical physics. Cambridge: Cambridge University Press.

    Google Scholar 

  • Baysan, U., & Wilson, J. (2017). Must strong emergence collapse? Philosophica, 91, 49–104.

    Google Scholar 

  • Bechtel, W. (2006). Discovering cell mechanisms: The creation of modern cell biology. Cambridge: Cambridge University Press.

    Google Scholar 

  • Bechtel, W. (2011). Mechanism and biological explanation. Philosophy of Science, 78(4), 533–557.

    Article  Google Scholar 

  • Bechtel, W. (2017). Explicating top-down causation using networks and dynamics. Philosophy of Science, 84(2), 253–274. https://doi.org/10.1086/690718.

    Article  Google Scholar 

  • Bechtel, W., & Abrahamsen, A. (2010). Dynamic mechanistic explanation: Computational modeling of circadian rhythms as an exemplar for cognitive science. Studies in History and Philosophy of Science, 41, 321–333.

    Article  Google Scholar 

  • Bechtel, W., & Richardson, R. (2010). Discovering complexity: Decomposition and localization as strategies in scientific research. Cambridge: The MIT Press.

    Book  Google Scholar 

  • Bedau, M. (1997). Weak emergence. In J. Tomberlin (Ed.), Philosophical perspectives: Mind, causation, and world (Vol. 11, pp. 375–399). Malden: Blackwell.

    Google Scholar 

  • Bitbol, M. (2012). Downward causation without foundations. Synthese, 185(2), 233–255.

    Article  Google Scholar 

  • Blanshard, B. (1962). Reason and analysis. London: George Allen & Unwin Ltd.

    Google Scholar 

  • Bohm, D. (1984). Causality and chance in modern physics. London: Routledge.

    Google Scholar 

  • Borghini, A. (2016). A critical introduction to the metaphysics of modality. London: Bloomsbury.

    Google Scholar 

  • Borghini, A., & Williams, N. (2008). A dispositional theory of possibility. Dialectica, 62, 21–41.

    Article  Google Scholar 

  • Broad, C. D. (1925). The mind and its place in nature. New York: Harcourt, Brace & Company.

    Google Scholar 

  • Clarke, R. (1999). Nonreductive physicalism and the causal powers of the mental. Erkenntnis, 51, 295–322.

    Article  Google Scholar 

  • Cohen, I., Atlan, H., & Efroni, S. (2016). Genetics as explanation: Limits to the human genome project. In K. E. Cullen (Ed.), Encyclopedia of life sciences (eLS). Chichester: Wiley. https://doi.org/10.1002/9780470015902.a0005881.pub3.

    Chapter  Google Scholar 

  • Craver, C., & Bechtel, W. (2007). Top-down causation without top-down causes. Biology and Philosophy, 22, 547–563.

    Article  Google Scholar 

  • Donati, P., & Archer, M. (2015). The relational subject. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Dunn, J. M. (1990a). Relevant predication 2: Intrinsic properties and internal relations. Philosophical Studies, 60(3), 177–206.

    Article  Google Scholar 

  • Dunn, J. M. (1990b). Relevant predication 3: Essential properties. In J. M. Dunn & A. Gupta (Eds.), Truth or consequences: Essays in Honor of Nuel Belnap (pp. 77–95). Dordrecht: Kluwer Academic Publishers.

    Chapter  Google Scholar 

  • Gamba, P., & Zenkin, N. (2018). Transcription fidelity and its roles in the cell. Current Opinion in Microbiology, 42, 13–18.

    Article  Google Scholar 

  • Ganeri, J. (2011). Emergence, ancient and modern. Mind, 120, 671–703.

    Article  Google Scholar 

  • Garson, J. (2006). Emergence. In S. Sarkar & J. Pfeifer (Eds.), The philosophy of science: An encyclopedia (pp. 230–235). New York: Routledge.

    Google Scholar 

  • Gillett, C. (2002). The dimensions of realization: A critique of the standard view. Analysis, 62, 316–323.

    Article  Google Scholar 

  • Gillett, C. (2016). Reduction and emergence in science and philosophy. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Glennan, S. (2017). The new mechanical philosophy. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Glennan, S., & Illari, P. (2018). Introduction: Mechanisms and mechanical philosophies. In S. Glennan & P. Illari (Eds.), The Routledge handbook of mechanisms and mechanical philosophy (pp. 1–9). New York: Routledge.

    Google Scholar 

  • Grelling, K., & Oppenheim, P. (1988). Logical analysis of ‘Gestalt’ as ‘functional whole’ [1939]. In B. Smith (Ed.), Foundations of Gestalt theory (pp. 210–217). Munich: Philosophia Verlag.

    Google Scholar 

  • Griffiths, A., Wessler, S., Lewontin, R., & Carroll, S. (2008). Introduction to genetic analysis (9th ed.). New York: W. H. Freeman and Company.

    Google Scholar 

  • Guay, A., & Sartenaer, O. (2016). A new look at emergence. Or when After is different. European Journal for Philosophy of Science, 6(2), 297–322.

    Article  Google Scholar 

  • Hempel, C. (1951). General system theory and the unity of science. Human Biology, 23, 313–327.

    Google Scholar 

  • Hempel, C. (1965). Aspects of scientific explanation and other essays in the philosophy of science. New York: The Free Press.

    Google Scholar 

  • Hendry, R. F., Gibb, S., & Lancaster, T. (2019). Introduction. In S. Gibb, R. Hendry, & T. Lancaster (Eds.), The Routledge handbook of emergence (pp. 1–19). London: Routledge.

    Google Scholar 

  • Humphreys, P. (2006). Emergence. In D. Borchert (Ed.), The encyclopedia of philosophy (Vol. 3, pp. 190–194). New York: Macmillan.

    Google Scholar 

  • Humphreys, P. (2016). Emergence. A philosophical account. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Hüttemann, A. (2004). What’s wrong with microphysicalism?. London: Routledge.

    Book  Google Scholar 

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

    Book  Google Scholar 

  • Kaplan, D. (2018). Mechanisms and dynamical systems. In S. Glennan & P. Illari (Eds.), The Routledge handbook of mechanisms and mechanical philosophy (pp. 267–280). New York: Routledge.

    Google Scholar 

  • Keller, E. (1995). Refiguring life. Metaphors of twentieth-century biology. New York: Columbia University Press.

    Book  Google Scholar 

  • Keller, E. (2001). Beyond the gene but beneath the skin. In S. Oyama, P. Griffiths, & R. Gray (Eds.), Cycles of contingency: Developmental systems and evolution (pp. 299–312). Cambridge: MIT Press.

    Google Scholar 

  • Keller, E. (2010). It is possible to reduce biological explanations in chemistry and/or physics? In F. J. Ayala & R. Arp (Eds.), Contemporary debates in philosophy of biology (pp. 19–31). Malden: Wiley-Blackwell.

    Google Scholar 

  • Kim, J. (1993). Supervenience and mind. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Kim, J. (1998). Mind in a physical world. Cambridge: The MIT Press.

    Book  Google Scholar 

  • Kim, J. (2006). Emergence: Core ideas and issues. Synthese, 151(3), 547–559.

    Article  Google Scholar 

  • Klee, R. (1984). Micro-determinism and concepts of emergence. Philosophy of Science, 51, 44–63.

    Article  Google Scholar 

  • Levins, R. (2011). Complex systems [1970]. In C. H. Waddington (Ed.), Organization, stability & process: Towards a theoretical biology (Vol. 3, pp. 73–87). New York: Routledge.

    Google Scholar 

  • Lewis, D. (1986). Philosophical papers (Vol. 2). Oxford: Oxford University Press.

    Google Scholar 

  • Lewis, D. (1994). Humean supervenience debugged. Mind, 103, 473–490.

    Article  Google Scholar 

  • Lewontin, R. (2000). The triple helix: Gene, organism, and environment. Cambridge: Harvard University Press.

    Google Scholar 

  • Lewontin, R. (2008). Heredity and heritability. In S. Sarkar & A. Plutynski (Eds.), A companion to the philosophy of biology (pp. 40–57). Malden: Blackwell.

    Google Scholar 

  • Lewontin, R., & Levins, R. (2007). Biology under the influence: Dialectical essays on ecology, agriculture and health. New York: Monthly Review Press.

    Google Scholar 

  • Mahner, M., & Bunge, M. (1997). Foundations of biophilosophy. Berlin: Springer.

    Book  Google Scholar 

  • Mill, J. S. (1974). A system of logic ratiocinative and inductive, Books I–III [1843]. In J. M. Robson (Ed.), Collected works of John Stuart Mill (Vol. 7). Toronto: University of Toronto Press.

    Google Scholar 

  • Moreno, A., & Mossio, M. (2015). Biological autonomy: A philosophical and theoretical inquiry. Dordrecht: Springer.

    Book  Google Scholar 

  • Morgan, L. (1923). Emergent evolution. London: Williams and Norgate.

    Google Scholar 

  • Nagel, E. (1951). Mechanistic explanation and organismic biology. Philosophy and Phenomenological Research, 11, 327–338.

    Article  Google Scholar 

  • Nagel, E. (1952). Wholes, sums, and organic unities. Philosophical Studies, 3(2), 17–32.

    Article  Google Scholar 

  • Nagel, E. (1961). Structure of science: Problems in the logic and scientific explanation. New York: Harcourt, Brace & World Inc.

    Book  Google Scholar 

  • O’Connor, T., & Wong, H. Y. (2005). The metaphysics of emergence. Noûs, 39(4), 658–678.

    Article  Google Scholar 

  • Oppenheim, P., & Putnam, H. (1958). Unity of science as a working hypothesis. In H. Feigl, M. Scriven, & G. Maxwell (Eds.), Concepts, theories, and the mind-body problem (pp. 3–36). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Pettit, P. (1993). A Definition of Physicalism. Analysis, 53, 213–223.

    Article  Google Scholar 

  • Povich, M., & Craver, C. (2018). Mechanistic levels, reduction, and emergence. In S. Glennan & P. Illari (Eds.), The Routledge handbook of mechanisms and mechanical philosophy (pp. 185–197). New York: Routledge.

    Google Scholar 

  • Quine, W. (2008). The way the world is [1986]. In D. Føllesdal & D. Quine (Eds.), Confessions of a confirmed extensionalist and other essays, pp 166–171. Cambridge: Harvard University Press.

    Chapter  Google Scholar 

  • Rescher, N., & Oppenheim, P. (1955). Logical analysis of gestalt concepts. The British Journal for the Philosophy of Science, 6(22), 89–106.

    Article  Google Scholar 

  • Santos, G. (2015a). Ontological emergence: How is that possible? Towards a new relational ontology. Foundations of Science, 20(4), 429–446.

    Article  Google Scholar 

  • Santos, G. (2015b). Upward and downward causation from a relational-horizontal ontological perspective. Axiomathes, 25(1), 23–40.

    Google Scholar 

  • Sartenaer, O. (2018). Flat emergence. Pacific Philosophical Quarterly, 99(1), 225–250.

    Article  Google Scholar 

  • Schaffer, J. (2010). The internal relatedness of all things. Mind, 119, 341–376.

    Article  Google Scholar 

  • Shapiro, J. (2009). Revisiting the central dogma in the 21st century. Annals of the New York Academy of Sciences, 1178, 6–28.

    Article  Google Scholar 

  • Shoemaker, S. (2002). Kim on emergence. Philosophical Studies, 108, 53-63.

    Article  Google Scholar 

  • Silberstein, M. (2017). Strong emergence no, contextual emergence yes. Philosophica, 91, 145–192.

    Google Scholar 

  • Silberstein, M. (2020). Constraints on localization and decomposition as explanatory strategies in the biological sciences 2.0. In F. Calzavarini & Viola, M. (Eds.), Neural mechanisms: New challenges in the philosophy of neuroscience. Springer Neuroscience (forthcoming).

  • Silberstein, M., & Chemero, A. (2013). Constraints on localization and decomposition as explanatory strategies in the biological sciences. Philosophy of Science, 80(5), 958–970.

    Article  Google Scholar 

  • Silberstein, M., & McGeever, J. (1999). The search for ontological emergence. The Philosophical Quarterly, 49(195), 182–200.

    Article  Google Scholar 

  • Simon, H. (1962). The architecture of complexity. Proceedings of the American Philosophical Society, 106(6), 467–482.

    Google Scholar 

  • Simons, P. (1987). Parts. A study in ontology. New York: Oxford.

    Google Scholar 

  • Simons, P. (2006). Real wholes, real parts: Mereology without algebra. The Journal of Philosophy, 103, 597–613.

    Article  Google Scholar 

  • Strohman, R. (1997). Epigenesis and complexity: The coming Kuhnian revolution in biology. Nature Biotechnology, 15, 194–200.

    Article  Google Scholar 

  • Vecchi, D. (2019). DNA is not an ontologically distinctive developmental cause. Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences. https://doi.org/10.1016/j.shpsc.2019.101245.

    Article  Google Scholar 

  • Wilson, J. (2015). Metaphysical emergence: Weak and strong. In T. Bigaj & C. Wüthrich (Eds.), Metaphysics in contemporary physics (pp. 345–402). Leiden: Brill-Rodo-pi.

    Google Scholar 

  • Wimsatt, W. (2006). Aggregate, composed, and evolved systems: Reductionistic heuristics as means to more holistic theories. Biology and Philosophy, 21, 667–702.

    Article  Google Scholar 

  • Wimsatt, W. (2007). Re-engineering philosophy for limited beings: Piecewise approximations to reality. Cambridge: Harvard University Press.

    Book  Google Scholar 

  • Winning, J., & Bechtel, W. (2018). Rethinking causality in biological and neural mechanisms: Constraints and control. Minds and Machines, 28(2), 287–310.

    Article  Google Scholar 

Download references

Acknowledgements

Thanks to the participants of the Cologne Conference on Diachronic Emergence, especially Olivier Sartenaer and Andreas Hüttemann for organising it. Thanks to Davide Vecchi for his helpful and stimulating comments on earlier versions of this paper. Finally, thanks to the referees for their insightful and constructive comments and suggestions. I acknowledge the financial support of FCT, ‘Fundação para a Ciência e a Tecnologia, I.P.’ (Stimulus of Scientific Employment, Individual Support 2017: CEECIND/03316/2017). This work is part of the FCT Project ‘Emergence in the Natural Sciences: Towards a New Paradigm’ (PTDC/FERHFC/30665/2017).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gil Santos.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Santos, G. Integrated-structure emergence and its mechanistic explanation. Synthese 198, 8687–8711 (2021). https://doi.org/10.1007/s11229-020-02594-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-020-02594-3

Keywords

Navigation