Skip to main content
Log in

A New Look at ‘Levels of Organization’ in Biology

  • Original Research
  • Published:
Erkenntnis Aims and scope Submit manuscript

[T]here is merit to the claim that much problem solving effort is directed at structuring problems, and only a fraction of it at solving problems once they are structured.

Herbert Simon 1977, 310.

Abstract

Despite its pervasiveness, the concept of ‘levels of organization’ has received relatively little attention in its own right. I propose here an emerging approach that posits ‘levels’ as a fragmentary concept situated within an interest-relative matrix of operational usage within scientific practice. To this end I propose one important component of meaning, namely the epistemic goal (sensu Ingo Brigandt) motivating the term’s usage, which recovers a remarkably conserved and sufficiently unifying significance attributable to ‘levels’ across different instances of usage. This epistemic goal, to provide structure to scientific problems, delegates tasks whose execution generates the term’s expressed content in a given instance. This treatment of levels does not diminish the concept’s general importance to science, but rather allows for its use in, and usefulness for, scientific practice to be better contextualized to particular tasks encompassing varying breadths of activity.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. Hereafter simply ‘levels’, unless otherwise specified; I assume here but will not argue for now, that ‘levels of organization’ possesses its own distinct identity among other “level” labels in that (i) ‘levels’ possesses definite cognitive boundaries (which I touch upon in Sect. 2 with respect to stratification, complexity and organization), (ii) ‘levels’ has its own distinct history in science, and (iii) ‘levels’ comes with its own distinct challenges.

  2. I do believe, however, that these questions are deeply connected, and will return to some upshots for a conception of the nature of levels in the conclusion.

  3. Before proceeding, I should specify two caveats: I do not mean to claim (i) that this is the only EG of the levels concept, nor (ii) that the EG of structuring scientific problems is held solely by the levels concept. The first caveat should be rather straightforward. Concerning the second, it is clear that this EG is a generic motivation surely shared or possessed by other scientific concepts. This for some may threaten the putative centrality of the levels concept in science and philosophy within its domain of application. I offer two reasons to downplay this threat. First, I believe we can be fully content with ‘levels’ sharing this EG with other concepts: ‘Levels’ is, after all and among other things, one among many organizing principles in the natural sciences. Second, and in this spirit, recovering any stable significance to the levels concept should count as progress towards excavating its widespread influences in the scientific literature. I will argue below, after all, that this EG can be shown to be extremely conserved across different uses of the levels concept.

  4. Using this basic distinction, Craver suggests we start thinking of “levels” as “primarily features of the world rather than as features of the units or products of science”, and that thinking of “levels” in terms of “sciences and theories” is “derivative upon” these ontological structures (2007, 177, fn; emphasis added). I cautiously agree with this sentiment, but strongly conditionalize my approach to ‘levels’ by separating what scientists express when using the term from what status the content expressed by ‘levels’ possesses. In this way I wish to explicitly make room for considering the significance of ‘levels’ in science for instances (a) where ‘levels’ refers primarily to epistemic constituents, or (b) where ‘levels’ does not in fact successfully refer to “ontic structures” in nature. My agreement with Craver’s sentiment is more along the lines that when scientists use ‘levels’, they take it to mean that nature recovers at least some of the content we express with levels, though this doesn’t have to be the case.

  5. This perspective does not exclude more ontologically-oriented questions (“What are levels of organization?”), and in fact arguably clears the way for a more adequate answer into such questions. I return to this in the conclusion.

  6. EGs may themselves be a source of variation, particularly when a concept is used to pursue multiple EGs, or when an EG serves multiple concepts. An EG provides coherence to a variable concept insofar as a stable EG (or multiple EGs) can be connected to a particular concept. As Brigandt explains: “While a scientific community often uses several concepts and theoretical resources to pursue a particular explanatory or investigative goal, in some cases such an epistemic goal (or a set of epistemic goals) can be tied to an individual scientific concept, in that the rationale for the introduction or continued use of a central theoretical concept is to pursue this epistemic goal” (Brigandt 2010, 23). The EG of ‘levels’ exhibits just such stability.

  7. To be sure, a problem agenda is a more complicated type of problem entity that subsumes the former two problem tasks of characterizing a problem and articulating constraints of the problem and its solution (as an anonymous reviewer pointed out). However, I take a problem agenda to also be (due to its more molar character) a more ‘organic’ kind of problem entity that subsumes or presupposes many more non-problem factors as part of its framework, such as social and research-political factors involved in the conceptual and empirical negotiations that adjudicate between the interests of the different scientific actors being served by the agenda. Thus, the contrast between the first two kinds of tasks and the task(s) involved in generating a problem agenda lend examples to different grains of problem-engaging behavior by researchers. More particularly, they encompass different breadths of scientific activity in which usage of ‘levels’ finds fertile ground.

  8. In this way, the so-called “top-down” approach was itself retooled from earlier meanings related to first looking to mathematically-derived “theoretical” descriptions of motion-mediated behaviors in flies (particularly the optomotor reflex) using abstract cybernetic models, then connecting these descriptions to physiological details that implement these algorithmic descriptions. As these passages here indicate, “top-down” now came to mean looking ‘deeper’ into the physiological details of the fly’s visual system components themselves in a level-mediated sense. To be sure, looking into how mathematical descriptions are realized in the fly brain is still an ongoing project, but this project has largely been displaced by “wet” biological studies into how components of the visual system actually process motion.

  9. One important source I have already intimated toward for these issues (Brooks 2017, 153) is William C. Wimsatt’s pioneering work on levels, which offers a potent and suggestive springboard into thinking about the nature of levels (see also Eronen and Brooks 2018). However, Wimsatt’s work is not without its attendant challenges. For example: What are “local maxima of regularity and predictability” (Wimsatt’s tentative definition for levels)? How do perspectives (or “causal thickets”) contrast with levels, in the Wimsattian sense or otherwise (see especially Kästner 2018)? Solutions to these questions comprise an early to-do list for any work into the nature of levels of organization.

  10. In another paper (Brooks, forthcoming), I offer an account of the “fragmentary” character of the levels concept, which traces the variation of the levels concept to four putative “content fragments”, or unique core attributes that each specify part of the meaning of ‘levels’ in a given instance.

  11. I am open to the possibility, though am not convinced, that the levels concept could turn out to be, as the skeptics claim, a hopelessly muddled concept that produces systematically misleading or false descriptive content. Though not probable (if I am right), there is certainly no ‘silver bullet’ aspect to the levels concept in the matters of its application. Particularly, I credit the work of Angela Potochnik and Markus Eronen on levels for compelling me to acknowledge that substantial influence of, or attributed to, ‘levels’ has been noxious or misleading. This, however, is to be expected of heuristic notions, which typically eschew principled accounts of scientific usefulness.

References

  • Bechtel, W., & Abrahamsen, A. (2005). Explanation: A mechanist alternative. Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences, 36, 421–441.

    Article  Google Scholar 

  • Beck, M., et al. (2011). Exploring the spatial and temporal organization of a cell’s proteome. Journal of Structural Biology, 173, 483–496.

    Article  Google Scholar 

  • Bickle, J. (2006). Reducing mind to molecular pathways: Explicating the reductionism implicit in current cellular and molecular neuroscience. Synthese, 151, 411–443.

    Article  Google Scholar 

  • Borst, A., & Egelhaaf, E. (1989). Principles of visual motion detection. Trends in Neuroscience, 12(8), 297–306.

    Article  Google Scholar 

  • Borst, A., & Haag, J. (2002). Neural networks in the cockpit of the fly. Journal of Comparative Physiology A, 188, 419–437.

    Article  Google Scholar 

  • Brigandt I. (2006). A theory of conceptual advance. Explaining conceptual change in evolutionary, molecular, and evolutionary developmental biology. Dissertation, University of Pittsburgh. Retrieved May 1, 2018 from http://etd.library.pitt.edu/ETD/available/etd-08032006–145211.

  • Brigandt, I. (2010). The epistemic goal of a concept: Accounting for the rationality of semantic change and variation. Synthese, 177(1), 19–40.

    Article  Google Scholar 

  • Brigandt, I. (2011). Natural kinds and concepts: A pragmatist and methodologically naturalistic account. In J. Knowles & H. Rydenfelt (Eds.), Pragmatism, science and naturalism (pp. 171–196). Frankfurt am Main: Peter Lang Publishing.

    Google Scholar 

  • Brigandt, I. (2012). The dynamics of scientific concepts: The relevance of epistemic aims and values. In U. Feest & F. Steinle (Eds.), Scientific concepts and investigative practice (pp. 75–104). Berlin: De Gruyter.

    Google Scholar 

  • Brigandt, I. (2015). From developmental constraint to evolvability: How concepts figure in explanation and disciplinary identity. In A. Love (Ed.), Conceptual change in biology (pp. 305–325). Boston: Boston Studies in the Philosophy and History of Science.

    Google Scholar 

  • Brigandt, I., & Love, A. C. (2012). Conceptualizing evolutionary novelty: moving beyond definitional debates. Journal of Experimental Zoology Part B: Molecular and Developmental Evolution., 318B, 417–427.

    Article  Google Scholar 

  • Brooks, D. S. (2014). The role of models in the process of epistemic integration: the case of the Reichardt motion detector. History and Philosophy of the Life Sciences, 36(1), 90–113.

    Article  Google Scholar 

  • Brooks, D. S. (2017). In defense of levels: Layer cakes and guilt by association. Biological Theory, 12(3), 142–156.

    Article  Google Scholar 

  • Brooks, D. S. (forthcoming). ‘Levels of organization’ as tool and doctrine in biology. In D. S. Brooks, J. DiFrisco, & W. C. Wimsatt (Eds.), Hierarchy and levels of organization in the biological sciences. MIT Press: Cambridge.

  • Brooks, D. S., & Eronen, M. I. (2018). The significance of ‘levels of organization’ for scientific research: A heuristic approach. Studies in History and Philosophy of Biological and Biomedical Sciences, 68, 34–41.

    Article  Google Scholar 

  • Bunge, M. (1960). Levels: A semantical preliminary. The Review of Metaphysics., 3(3), 396–406.

    Google Scholar 

  • Craver, C. F. (2001). Role functions, mechanisms, and hierarchy. Philosophy of Science, 68, 53–74.

    Article  Google Scholar 

  • Craver, C. F. (2007). Explaining the Brain. Mechanisms and the mosaic unity of neuroscience. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Craver, C. F. (2015). Levels. In T. Metzinger & J. M. Windt (Eds.), Open MIND (Vol. 26, pp. 1–26). Frankfurt am Main: MIND Group.

    Google Scholar 

  • Darden, L. (1991). Theory change in science: Strategies from Mendelian genetics. Oxford: Oxford University Press.

    Google Scholar 

  • DiFrisco, J. (2017). Time scales and levels of organization. Erkenntnis, 82(4), 795–818.

    Article  Google Scholar 

  • Douglass, J. K., & Strausfeld, N. (1996). Visual motion-detection circuits in flies: Parallel direction- and non-direction-sensitive pathways between the Medulla and Lobula Plate. The Journal of Neuroscience, 16(15), 4551–4562.

    Article  Google Scholar 

  • Edel, A. (1988). Integrative levels: Some reflections on a philosophical dimension. In G. Greenberg & E. Tobach (Eds.), Evolution of social behavior and integrative levels (pp. 65–74). New Jersey: Lawrence Erlbaum Associates.

    Google Scholar 

  • Egelhaaf, M., & Borst, A. (1993). A look into the cockpit of the fly: Visual orientation, algorithms, and identified neurons. Journal of Neuroscience, 13(11), 4563–4574.

    Article  Google Scholar 

  • Eldredge, N., et al. (Eds.). (2016). Evolutionary theory: A hierarchical perspective. Chicago: Chicago University Press.

    Google Scholar 

  • Engreitz, J. M., et al. (2016). Long non-coding RNAs: Spatial amplifiers that control nuclear structure and gene expression. Nature Reviews Molecular Cell Biology, 17, 756–770.

    Article  Google Scholar 

  • Eronen, M. I. (2013). No levels, no problems: Downward causation in neuroscience. Philosophy of Science, 80(5), 1042–1052.

    Article  Google Scholar 

  • Eronen, M. I. (2015a). Levels of organization: A deflationary account. Biology and Philosophy, 30(1), 39–58.

    Article  Google Scholar 

  • Eronen, M. I. (2015b). Robustness and reality. Synthese, 192(12), 3961–3977.

    Article  Google Scholar 

  • Eronen, M. I., & Brooks, D. S. (2018). Levels of organization in biology. In E. N. Zalta (Ed.), The stanford encyclopedia of philosophy (Spring 2018 edn). https://plato.stanford.edu/archives/spr2018/entries/levels-org-biology/.

  • Giere, R. N. (1988). Explaining science: A cognitive approach. Chicago: Chicago University Press.

    Book  Google Scholar 

  • Green, S. (forthcoming). Cancer beyond genetics: On the practical implications of downward causation. In D. S. Brooks, J. DiFrisco, & W. C. Wimsatt (Eds.), Hierarchy and levels of organization in the biological sciences. MIT Press: Cambridge.

  • Grene, M. (1969). Hierarchy: One word, how many concepts? In L. L. Whyte, et al. (Eds.), Hierarchical structures (pp. 56–58). New York: American Elsevier Publishing Company.

    Google Scholar 

  • Grene, M. (1987). Hierarchies in biology. American Scientist, 75(5), 504–510.

    Google Scholar 

  • Grillner, S., et al. (2005). Integrative neuroscience: linking levels of analyses. Current Opinion in Neurobiology, 15, 614–621.

    Article  Google Scholar 

  • Kaiser, M. I. (2015). Reductive Explanation in the Biological Sciences. History, Philosophy and Theory of the Life Sciences: (Vol. 16). Utrecht: Springer.

    Google Scholar 

  • Kästner, L. (2018). Integrating mechanistic explanations through epistemic perspectives. Studies in History and Philosophy of Science Part A, 68, 68–79.

    Article  Google Scholar 

  • List C. (2016). Levels: Descriptive, explanatory, and ontological. [Preprint]. PhilSci Archive.

  • Love, A. C. (2008). Explaining evolutionary innovations and novelties: criteria of explanatory adequacy and epistemological prerequisites. Philosophy of Science, 75, 874–886.

    Article  Google Scholar 

  • Mader, Sylvia S. (2010). Biology (10th ed.). McGraw Higher Education: Boston.

    Google Scholar 

  • Nersessian, N. (1992). How do scientists think? Capturing dynamics of conceptual change in science. In R. Giere & H. Feigl (Eds.), Cognitive models of science (pp. 3–45). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Nickles, T. (1978). Scientific problems and constraints. Proceedings of Philosophy of Science, 1978(1), 134–148.

    Google Scholar 

  • Nickles, T. (1981). What is a problem that we may solve it? Synthese, 47, 85–118.

    Article  Google Scholar 

  • Pavé, A. (2006). Hierarchical organization of biological and ecological systems. In D. Pumain (Ed.), Hierarchy in Natural and Social Sciences (pp. 39–70). Dordrecht: Springer.

    Chapter  Google Scholar 

  • Potochnik A. (forthcoming). Our world isn’t organized into levels. In D. S. Brooks, J. DiFrisco, & W. C. Wimsatt (Eds.) Hierarchy and levels of organization in the biological sciences. MIT Press: Cambridge.

  • Potochnik, A., & McGill, B. (2012). The limitations of hierarchical organization. Philosophy of Science, 79, 120–140.

    Article  Google Scholar 

  • Salthe, S. (1988). Notes toward a formal history of the levels concept. In G. Greenberg & E. Tobach (Eds.), Evolution of social behavior and integrative levels (pp. 53–64). New Jersey: Lawrence Erlbaum Associates.

    Google Scholar 

  • Simon, H.A. (1977). The structure of ill-structured problems. In Models of discovery and other topics in the methods of science (pp. 304–25). Dordrecht: Reidel.

  • Simon, H. A., & Newell, A. (1970). Human problem solving: The state of the theory in 1970. American Psychologist, 26(2), 145–159.

    Article  Google Scholar 

  • Steinle, F. (2012). Goals and fates of concepts: The case of magnetic poles. In U. Feest & F. Steinle (Eds.), Scientific concepts and investigative practice (pp. 105–126). Berlin: De Gruyter.

    Google Scholar 

  • Stieve, H. (1998). Four levels of brain research—An introduction. In A. Neugebauer (Ed.), Macromolecular Interplay in Brain Associative Mechanisms (pp. 309–319). Singapore: World Scientific.

    Chapter  Google Scholar 

  • Thalos, M. (2013). Without hierarchy: The scale freedom of the universe. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Umerez, J. (2016). Biological organization from a hierarchical perspective: Articulation of concepts and interlevel relation. In N. Eldredge, et al. (Eds.), Evolutionary theory: A hierarchical perspective (pp. 63–85). Chicago: Chicago University Press.

    Google Scholar 

  • Urry, L. A., et al. (2017). Campbell biology (11th ed.). San Francisco: Benjamin Cummings.

    Google Scholar 

  • Wimsatt, W. C. (1976). Reductionism, levels of organization, and the mind-body problem. In G. G. Globus, et al. (Eds.), Consciousness and the brain (pp. 205–267). New York: Plenum Press.

    Chapter  Google Scholar 

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

    Book  Google Scholar 

  • Winther, R. G. (2011). Part-whole science. Synthese, 178, 397–427.

    Article  Google Scholar 

  • Woodward, A. M. (1977). The roles of reviews in information transfer. Journal of the American Society for Information Science., 28(3), 175–180.

    Article  Google Scholar 

  • Woodward, J. (forthcoming). Downward causation and levels. In: D. S. Brooks, J. DiFrisco, & W. C. Wimsatt (Eds.) Hierarchy and levels of organization in the biological sciences. MIT Press: Cambridge.

  • Woody, A. (2003). On Explanatory Practice and Disciplinary Identity. Annals of the New York Academy of Sciences, 988, 22–29.

    Article  Google Scholar 

Download references

Acknowledgements

This paper benefitted substantially from my discussions with Alan Love, Wes Anderson, Ingo Brigandt, Markus Eronen, and James DiFrisco. I thank them all for their invaluable feedback and suggestions. I also benefited from the comments and queries of two anonymous reviewers; I thank them for their excellent suggestions and advice.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel S. Brooks.

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

Brooks, D.S. A New Look at ‘Levels of Organization’ in Biology. Erkenn 86, 1483–1508 (2021). https://doi.org/10.1007/s10670-019-00166-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10670-019-00166-7

Navigation