Skip to main content
Log in

Constitutive relevance & mutual manipulability revisited

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

A Correction to this article was published on 09 July 2021

This article has been updated

Abstract

An adequate understanding of the ubiquitous practice of mechanistic explanation requires an account of what Craver (J Philos Res, 32:3–20, 2007b) termed “constitutive relevance.” Entities or activities are constitutively relevant to a phenomenon when they are parts of the mechanism responsible for that phenomenon. Craver’s mutual manipulability (MM) account extended Woodward’s account of manipulationist counterfactuals to analyze how interlevel experiments establish constitutive relevance. Critics of MM (e.g., Baumgartner and Casini, Philos Sci 84:214–233, 2017; Baumgartner and Gebharter, Brit J Philos Sci 67:731–756, 2016) argue that applying Woodward’s account to this philosophical problem conflates causation and constitution, thus rendering the account incoherent. These criticisms, we argue, arise from failing to distinguish the semantic, epistemic, and metaphysical aspects of the problem of constitutive relevance. In distinguishing these aspects of the problem and responding to these critics accordingly, we amend MM into a refined epistemic criterion, the “matched interlevel experiments” (MIE) account. Further, we explain how this epistemological thesis is grounded in the plausible metaphysical thesis that constitutive relevance is causal betweenness.

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

Similar content being viewed by others

Change history

Notes

  1. We treat Baumgartner, Gebharter, and Casini’s as a single critical commentary, though we acknowledge that they might not all agree on every point.

  2. As Illari and Williamson (2012) note, the ontological significance of the many formulations of “mechanism” is often overstated. We use Glennan’s (2017) definition of “minimal mechanism”, which itself borrows MDC’s (2000) convention of distinguishing “entities,” the component parts, and “activities,” the component processes. Interactions are activities among two or more entities (Tabery, 2004). See Glennan (2017), Kaiser (2018), and Krickel (2018).

  3. One persistent criticism of the mechanistic approach is that it ignores the processual character of biological systems (Dupré, 2013; Dupré and Nicholson, 2018; Nicholson, 2012). This is puzzling given that processes were at the heart of Salmon’s philosophy and that more recent mechanistic approaches have often adopted that processual orientation explicitly (see, e.g., MDC, 2000; Glennan, 2017).

  4. Reflection on examples suggests that embodiment (in an S) comes in degrees. But to elucidate and defend this claim would require us to articulate more clearly the criteria by which entities are identified and individuated, and the conditions under which a collection of entities compose another entity. See Gillett (2016) and Glennan (2020) for recent discussions. Because the constitutive relevance problem is about the processual core of mechanisms, we bracket this question.

  5. Woodward (2020) notes that the value of the cause variable cannot in this case be fixed independently of the value of the would-be effect variable. This is true, however, precisely for the deeper reason that the two stand in a part-whole relationship to one another, as mechanists emphasize.

  6. To reject interlevel causal relations is consistent with emphasizing the import of higher-level causes in multilevel mechanisms. Mechanists generally (Craver 2007a, b; Glennan 2010, 2017) defend the existence and explanatory relevance of higher-level causes while rejecting the idea of interlevel causation as incoherent. Even if scientists do not routinely guard these metaphysical distinctions, they must be guarded nonetheless if we are to avoid speaking nonsense.

  7. Some (e.g., Baumgartner & Casini 2017 and two anonymous reviewers) interpret Craver as offering a necessary and sufficient condition, but the textual evidence against this interpretation seems unambiguous. Craver describes MM as a sufficient condition (2007a, pp. 104, 141, 159 twice; 2007b, p. 17). His formal specification is articulated using “if” (2007a, p. 154) or “when” (p. 153; 2007b, p. 15) rather than “only if,” “if and only if,” or “when and only when.” Finally, he offers counterexamples to show the thesis is not a necessary condition (2007a, beginning on p. 159; 2007b, p. 17) and so only a “suitable starting point for an account of constitutive relevance” (2007a, p. 160).

  8. Baumgartner and Cassini’s critique of MM rests in part on treating MM as a necessary and sufficient condition. Their positive argument for treating it as such despite the textual evidence in note 7, we believe, comes from a misreading of MM. They argue as follows:

    MM provides a sufficient condition for constitutive relevance and a sufficient condition for constitutive irrelevance, which jointly amount to a sufficient and necessary condition for constitutive relevance—that is, to a complete definition of constitution (Baumgartner & Casini 2017, p. 218).

    But their reasoning is fallacious. Consider the following formalization:

    R: X’s ϕ-ing is constitutively relevant to S’s ψ-ing (constitutive relevance)

    B: manipulating ϕ changes ψ (bottom-up manipulability)

    T: manipulating ψ changes ϕ (top-down manipulability)

    Baumgartner and Casini take Craver to assert the following sufficient conditions for relevance and irrelevance respectively:

    (1) (B & T) ⊃ R

    (2) ~(B & T) ⊃ ~R

    Together, (1) and (2) are indeed equivalent to a necessary and sufficient condition:

    (3) R ≡ (B & T)

    However, Craver does not assert (2), and it appears Baumgartner and Casini have misplaced the negation in (2). Craver’s actual statement of the sufficient condition is:

    To establish that a component is irrelevant, it is sufficient to show that one cannot manipulate S’s ψ-ing by intervening to change X’s ϕ-ing and that one cannot manipulate X’s ϕ-ing by manipulating S’s ψ-ing (Craver, 2007a, b, p. 159).

    Using our formalization:

    (2C) (~B & ~T) ⊃ ~R

    Crucially, 1 and 2C do not together entail a biconditional. As Craver writes, in cases where B is true but T is false, or vice versa, constitutive relevance is indeterminate.

  9. We do not doubt that wholes supervene on the organized collections of their parts, but the nature of this relation is a matter of dispute. We focus on the part-whole relation because it is better understood, and the putative challenge to MM comes directly from the impossibility of causal relations obtaining between parts and wholes.

  10. Parentheses indicate that ψ may be more or less embodied and so S is not essential to the parthood relation we consider.

  11. Some critics reject high-level interventions as “fat-handed”, a term borrowed from, e.g., Scheines (2005). A fat-handed intervention on ψ changes more variables than ψ, including possibly ϕ. But fat-handedness per se is not the problem here. Fat-handed interventions are problematic when the same intervention simultaneously manipulates two variables, each of which independently could be causally relevant de facto to the putative effect. In this case, however, the variables are not independent competitors. Rather, the intervention on the putative cause necessarily changes the putative effect. That is not de facto fat-handedness but rather a metaphysically necessary fat-handedness. Our two reconstructions of the problem here dispense with this obfuscatory ambiguity.

  12. Top-down inhibitory experiments are the control conditions for top-down excitatory experiments. Whether an intervention is “excitatory” and “inhibitory” depends on how something is situated in a causal context.

  13. Pearl (2012) generalizes the same idea to non-binary relations. That our reconstruction captures the idea of “causal mediation” developed in the causal modeling literature gives us some confidence that this integration of ontology and epistemology is on the right track. This convergence of qualitative discussions of mechanisms with quantitative work on causal modeling is a desired outcome, not a result to be shunned.

  14. Note that this reinterpretation is not entirely novel; Craver (2007a, b, pp. 145–146) warned against this confusion and hinted at this solution.

  15. Note that MIE drops the now-unnecessary parthood condition in Craver (2007a, b). If we take S simply to name the collection of all and only the Xs involved in some ψ-ing, the parthood condition is tautologous and uninformative, as Leuridan (2012) has urged. This is a side-benefit of our revision.

  16. Causal betweenness is a three-place relation, B(x, y, z), between events, and for some event, Xi’s ϕi-ing, to be constitutively relevant to a mechanism that ψs, it must be the case that B(ψin, Xi’s ϕi-ing, ψout). The Xi’s ϕi-ing shown in the lower half of Fig. 1 in fact lie causally between ψin and ψout (as shown in Fig. 2). Although the relata of the betweenness relation are events, we also speak of entities Xi or activities ϕi as lying between, when those entities and activities are constituents of the events that lie between. We understand events, in this context, as entities engaging in activities or interactions (see Glennan, 2017). Betweenness is thus fundamentally a relation between particular events. Further work is required to extend this analysis to types of events.

References

  • Aizawa, K., & Headley, D. (Forthcoming). Abduction and composition. Philosophy of Science.

  • Aizawa, K., & Gillett, C. (Eds.). (2016). Scientific composition and metaphysical ground. Palgrave-Macmillan.

    Google Scholar 

  • Aizawa, K., & Gillett, C. (2019). Defending pluralism about compositional explanations. 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.2019.101202

    Article  Google Scholar 

  • Anscombe, G. E. M. (1971). Causality and determination: An inaugural lecture. Cambridge: Cambridge University Press.

    Google Scholar 

  • Baumgartner, M., & Gebharter, A. (2016). Constitutive relevance, mutual manipulability, and fat-handedness. British Journal for the Philosophy of Science, 67(3), 731–756

    Google Scholar 

  • Baumgartner, M., & Casini, L. (2017). An abductive theory of constitution. Philosophy of Science, 84(2), 214–233

    Google Scholar 

  • Bechtel, W., & Richardson, R. C. (2010 [1993]). Discovering complexity: Decomposition and localization as strategies in scientific research. (2nd ed.). MIT Press/Bradford Books.

  • Bogen, J. (2008). Causally productive activities. Studies in History and Philosophy of Science Part A, 39(1), 112–123

    Google Scholar 

  • Couch, M. B. (2011). Mechanisms and constitutive relevance. Synthese, 183(3), 375–388. https://doi.org/10.1007/s11229-011-9882-z

    Article  Google Scholar 

  • Craver, C. F. (2001). Role functions, mechanisms, and hierarchy. Philosophy of Science, 68(1), 53–74

    Google Scholar 

  • Craver, C. F. (2005). Beyond reduction: Mechanisms, multifield integration and the unity of neuroscience. Studies in History and Philosophy of Science Part c: Studies in History and Philosophy of Biological and Biomedical Sciences, 36(2), 373–395

    Google Scholar 

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

    Google Scholar 

  • Craver, C. F. (2007b). Constitutive explanatory relevance. Journal of Philosophical Research, 32, 3–20.

    Google Scholar 

  • Craver, C. F., & Bechtel, W. (2007). Top-down causation without top-down causes. Biology & Philosophy, 22(4), 547–563

    Google Scholar 

  • Craver, C. F., & Darden, L. (2001). Discovering mechanisms in neurobiology: The case of spatial memory. In P. K. Machamer, R. Grush, & P. McLaughlin (Eds.), Theory and method in neuroscience. (pp. 112–137). University of Pittsburgh Press.

    Google Scholar 

  • Craver, C. F., & Darden, L. (2013). In search of mechanisms: Discoveries across the life sciences. University of Chicago Press.

    Google Scholar 

  • Darden, L. (2006). Reasoning in biological discoveries: Essays on mechanisms, interfield Relations, and anomaly resolution. Cambridge University Press.

    Google Scholar 

  • Dowe, P. (2000). Physical causation. Cambridge University Press.

    Google Scholar 

  • Dupré, J. (2013). Living causes. Aristotelian Society Supplementary, 87(1), 19–37

    Google Scholar 

  • Dupré, J., & Nicholson, D. J. (2018). A manifesto for a processual philosophy of biology. In D. J. Nicholson & J. Dupré (Eds.), Everything flows: Towards a processual philosophy of biology. (pp. 1–58). Oxford University Press.

    Google Scholar 

  • Ehring, D. (1997). Causation and persistence. Oxford University Press.

    Google Scholar 

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

    Google Scholar 

  • Gebharter, A. (2017). Uncovering constitutive relevance relations in mechanisms. Philosophical Studies, 174(11), 2645–2666

    Google Scholar 

  • Glennan, S. S. (1996). Mechanisms and the nature of causation. Erkenntnis, 44(1), 49–71

    Google Scholar 

  • Glennan, S. (2002). Rethinking mechanistic explanation. Philosophy of Science, 69(S3), S342–S353

    Google Scholar 

  • Glennan, S. S. (2010). Ephemeral mechanisms and historical explanation. Erkenntnis, 72(2), 251–266

    Google Scholar 

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

    Google Scholar 

  • Glennan, S. S. (2020). Corporeal composition. Synthese. https://doi.org/10.1007/s11229-020-02805-x

    Article  Google Scholar 

  • Harbecke, J. (2010). Mechanistic constitution in neurobiological explanations. International Studies in the Philosophy of Science, 24(3), 267–285

    Google Scholar 

  • Harinen, T. (2018). Mutual manipulability and causal inbetweenness. Synthese, 195(1), 35–54

    Google Scholar 

  • Illari, P. M., & Williamson, J. (2012). What is a mechanism? Thinking about mechanisms across the sciences. European Journal for Philosophy of Science, 2(1), 119–135

    Google Scholar 

  • Illari, P. M. (2011). Mechanistic evidence: Disambiguating the Russo-Williamson thesis. International Studies in the Philosophy of Science, 25(2), 139–157

    Google Scholar 

  • Kaiser, M. (2018). The components and boundaries of mechanisms. In S. Glennan & P. M. Illari (Eds.), The routledge handbook of mechanisms and mechanical philosophy. (pp. 116–130). Routledge.

    Google Scholar 

  • Kim, J. (2000). Making sense of downward causation. In P. B. Anderson, C. Emmeche, N. O. Finnemann, & P. V. Christiansen (Eds.), Downward causation. (pp. 305–321). Aarhus University Press.

    Google Scholar 

  • Krickel, B. (2018). The mechanical world: The metaphysical commitments of the new mechanistic approach. Springer.

    Google Scholar 

  • Ladyman, J., & Ross, D. (2007). Every thing must go: Metaphysics naturalized. Oxford University Press on Demand.

    Google Scholar 

  • LeDoux, J. (2003). The emotional brain, fear, and the amygdala. Cellular and Molecular Neurobiology, 23(4), 727–738

    Google Scholar 

  • Leuridan, B. (2012). Three problems for the mutual manipulability account of constitutive relevance in mechanisms. The British Journal for the Philosophy of Science, 63(2), 399–427

    Google Scholar 

  • Lewis, D. (2004). Causation as influence. In J. Collins, N. Hall, & L. Paul (Eds.), Causation and Counterfactuals (pp. 75–106). Cambridge, MA: Bradford Books, MIT Press.

    Google Scholar 

  • Machamer, P., Darden, L., & Craver, C. F. [MDC] (2000). Thinking about mechanisms. Philosophy of Science, 67(1), 1–25.

  • Mackie, J. L. (1974). The cement of the universe: A study of causation. Clarendon Press.

    Google Scholar 

  • Nicholson, D. J. (2012). The concept of mechanism in biology. Studies in History and Philosophy of Science Part c: Studies in History and Philosophy of Biological and Biomedical Sciences, 43(1), 152–163

    Google Scholar 

  • Pearl, J. (2012). The mediation Formula: A guide to the assessment of causal pathways in nonlinear models. In C. Berzuini, P. Dawid, & L. Bernardinell (Eds.), Causality: Statistical perspectives and applications. (pp. 151–179). Wiley.

    Google Scholar 

  • Prychitko, E. (2019). The causal situationist account of constitutive relevance. Synthese. https://doi.org/10.1007/s11229-019-02170-4

    Article  Google Scholar 

  • Reichenbach, H. (1956). The direction of time. University of California Press.

    Google Scholar 

  • Romero, F. (2015). Why there isn’t inter-level causation in mechanisms. Synthese, 192(11), 3731–3755

    Google Scholar 

  • Salmon, W. (1980). Probabilistic causality. Pacific Philosophical Quarterly, 61(1–2), 50–74

    Google Scholar 

  • Salmon, W. (1998). Causality and explanation. Oxford University Press. https://doi.org/10.1093/0195108647.001.0001

    Book  Google Scholar 

  • Salmon, W. C. (1984). Scientific explanation and the causal structure of the world. Princeton University Press.

    Google Scholar 

  • Scheines, R. (2005). The similarity of causal inference in experimental and non-experimental studies. Philosophy of Science, 72(5), 927–940

    Google Scholar 

  • Sengupta, P., & Samuel, A. D. (2009). Caenorhabditis elegans: A model system for systems neuroscience. Current Opinion in Neurobiology, 19(6), 637–643

    Google Scholar 

  • Spirtes, P., Glymour, C. N., Scheines, R., & Heckerman, D. (2000). Causation, prediction, and search. MIT Press.

    Google Scholar 

  • Stevens, C. F. (1998). A million dollar question: Does LTP = memory? Neuron, 20(1), 1–2

    Google Scholar 

  • Tabery, J. G. (2004). Synthesizing activities and interactions in the concept of a mechanism. Philosophy of Science, 71(1), 1–15

    Google Scholar 

  • Wilhelm, I. (2019). The ontology of mechanisms. The Journal of Philosophy, 116(11), 615–636

    Google Scholar 

  • Woodward, J. (2003). Making things happen: A theory of causal explanation. Oxford University Press.

    Google Scholar 

  • Woodward, J. (2020). Causal complexity, conditional independence, and downward causation. Philosophy of Science, 87(5), 857–867.

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mark Povich.

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

Craver, C.F., Glennan, S. & Povich, M. Constitutive relevance & mutual manipulability revisited. Synthese 199, 8807–8828 (2021). https://doi.org/10.1007/s11229-021-03183-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-021-03183-8

Keywords

Navigation