Skip to main content

Two Perspectives on Explanatory Reduction

  • Chapter
  • First Online:
Reductive Explanation in the Biological Sciences

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

  • 673 Accesses

Abstract

In Chapter 4 I critically discuss the two perspectives on explanatory reduction that have been proposed in the philosophy of biology so far, namely Rosenberg’s thesis that explanatory reduction is a relation between a higher-level and a lower-level explanation of the same phenomenon and Sarkar’s, Hüttemann’s, and Love’s approach to focus on individual reductive explanations. The result of my critical examination will be that Rosenberg’s perspective on explanatory reduction in biology has several shortcomings and that, even though Sarkar’s, Hüttemann’s, and Love’s encounters objections, too, it seems to be the more promising path to run.

Reductionism is the thesis that biologicalexplanationsneed to be grounded in molecular biology and ultimately physical science, for it is only by doing so that they can be improved, corrected, strengthened, made more accurate and more adequate, and completed.” (Alex Rosenberg 2006, 4)

“[W]e must ask what substantive criteria distinguish reductionist explanations from other forms of explanation. Reductionism then becomes the empirical thesis that explanations in a particular discipline satisfy those criteria.” (Sahotra Sarkar 2008, 427)

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 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
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.

    Also Weber evaluates this as a drawback when he claims that Rosenberg’s book “suffers a bit from the lack of an explicit account of reduction” (2008, 151).

  2. 2.

    To get an overview of these difficulties see, for example, Weber (2008) and Love (2008b).

  3. 3.

    Besides his three major books “The Structure of Biological Science” (1985), “Instrumental Biology or the Disunity of Science” (1994), and “Darwinian Reductionism” (2006) he has published an immense bulk of papers.

  4. 4.

    This deprecatory judgment about the limited cognitive powers of humans can also be found in Rosenberg’s latest work. But here it appears as a deficient argumentative strategy of the explanatory antireductionist (see 2006, 14f, 36).

  5. 5.

    Since Rosenberg also speaks about inadequate and false explanations he does not seem to use the term ‘explanation’ as a success term.

  6. 6.

    There is a possible objection to Rosenberg’s claim that maximally complete molecular explanations are always better or the only adequate explanations. For instance, Elliott Sober (1999) has argued that depth is only one of two possible virtues of explanations (the other is generality). This point is elaborated in Chap. 5, Sect. 4.4.

  7. 7.

    More about Rosenberg’s view on explanation can be found in Chap. 5.

  8. 8.

    I am aware of the fact that Rosenberg does not use the term ‘explanatory reduction’. However, he is concerned with reduction as a relation that exists between explanations. Hence, there is no reason why one should not classify this kind of reduction as explanatory reduction (together with individual reductive explanations, which will be discussed in Sect. 2).

  9. 9.

    For the sake of simplicity I will sometimes also talk about higher-level explanations being reduced to lower-level explanations. However, I am aware of the fact that the distinction between higher-level and lower-level explanations does not completely coincide with the distinction between ultimate and proximate explanations.

  10. 10.

    At this point one might doubt whether this “adding molecular details to an ultimate explanation” process really can yield a proximate explanation, as Rosenberg claims. I agree that this is a problematic step in Rosenberg’s argumentation. However, one should note that Rosenberg states that “the reductionist’s full explanation is still a historical explanation in which further historical facts – about genes and pathways – are added, and are connected by the same principles of natural selection that are invoked by the ultimate functional how-possible explanation” (2006, 46). But, in my view, this statement does not clarify much. Even worse, it completely blurs the difference between ultimate and proximate explanations.

  11. 11.

    In Rosenberg’s view there exists only one genuine biological law, namely the principle of natural selection, which is responsible for the explanatory force of all biological explanations (see 2006, Chapter 4).

  12. 12.

    This claim is supported by the various examples of paradigmatic and important cases of explanatory reduction I discuss in Chap. 6.

  13. 13.

    Love calls this an “artificial ‘two-levellism’” (2008b, 3).

  14. 14.

    The close link between the question of reduction(ism) and the question of explanation does not imply that developing an account of explanatory reduction amounts to nothing more than developing an account of explanation. The contrary is the case, as I will elaborate in Chap. 6.

  15. 15.

    The empirical evidence for this claim can be found in Chap. 6.

  16. 16.

    For instance Wimsatt writes: “At least in biology, most scientists see their work as explaining types of phenomena by discovering mechanisms… and this is seen by them as reduction” (1976a, 671).

  17. 17.

    Wimsatt’s paper “Reductive Explanation: A Functional Account” (1976a) is reprinted as a slightly modified version in his newest book “Re-engineering Philosophy for Limited Beings” (2007). In what follows I cite only the original paper.

  18. 18.

    The great interest in the discovery of explanations is representative for the New Mechanistic Philosophy, too (see, in particular, Craver and Darden 2001, 2005; Craver 2002b; Bechtel 2006).

  19. 19.

    It is interesting and in a way surprising that many of the ideas of the New Mechanists, which at first sight appear to be brand-new, already can be found in Kauffman’s paper.

  20. 20.

    However, such an attempt is problematic in two respects: on the one hand, Nagel’s notion of non-formal conditions for reduction does not concur with the non-formal character of Sarkar’s or Wimsatt’s account. According to Nagel, the additional non-formal conditions for reduction serve to distinguish trivial from noteworthy scientific achievements since they require that theoretical assumptions must be supported by empirical evidence and must exhibit some degree of predictive power (1961, 358–361). This is not the same as what Sarkar and Wimsatt have in mind when they speak about the non-formal character of their account of reduction. On the other hand, Nagel’s non-formal conditions do not suffice on their own to distinguish cases of epistemic reductions from non-reductions. This is due to the fact that they hold for all theoretical assumptions in science. Hence, there is a good reason why Sarkar and Wimsatt do not adopt Nagel’s notion of non-formality: his non-formal conditions for reduction fail to provide a basis for an adequate non-formal analysis of epistemic reduction.

  21. 21.

    This is in line with Sarkar’s other remarks on the substantive character of his account. For instance, he refers to his criteria of reductivity as ‘substantive’ because “they are about what assumptions are made during a (putative) reductionist explanation, rather than about the form that such an explanation may take” (1998, 43).

  22. 22.

    Objects on different spatial levels are related to each other via part-whole relations or relations of spatial containment, e.g., parts of a system are located on a lower spatial level than the system (Sarkar 1998, 55).

  23. 23.

    Sarkar argues that abstract hierarchies need to be “constructed according to some independent criterion” (1998, 43). This condition serves to preclude rare cases, in which the hierarchy is postulated only for the sake of the explanation.

  24. 24.

    According to Sarkar it is irrelevant that at least a part of the allele-genotype-phenotype hierarchy could in principle be spelled out spatially (since genomes spatially consist of linkage groups, which in turn consist of loci that are occupied by alleles). What is important is not whether the hierarchy might to some extent be spelled out spatially, but that the hierarchy is in fact represented as an abstract and not as a spatial hierarchy in reductive explanations in genetics (1998, 128).

  25. 25.

    It is important to note that this characterization of the reductivity of genetic explanation does not imply the admittedly controversial thesis that reductions of this kind will always be successful. Particularly, in the case of mental disease traits or other complex behavioral traits the development of a genetic explanation is highly problematic if not impossible. For a detailed discussion of these problems compare Sarkar 1998, 124–127, 131f.

  26. 26.

    In a later paper Sarkar refers to this criterion also as the “criterion of compositionality” (2008, 429).

  27. 27.

    A notable exception constitutes Sarkar’s philosophical work on ecology (e.g., Sarkar 2005, 111f; 2009).

  28. 28.

    In a recent paper Sarkar seems to recognize this drawback (at least with respect to research strategies): “All potentially non-reductionist research strategies discussed satisfy the criterion of epistemic primacy required for weak reductions.” (2008, 433)

  29. 29.

    One might object that characterizing a factor or a kind of factors (or a realm) as fundamental involves more than just characterizing it as explanatorily relevant or prior. This is true for our intuitive understanding of the term ‘fundamental’. However, the way Sarkar spells out his “fundamentalism” criterion lacks this supplement. He leaves it almost completely unspecified what it means for one realm to be more fundamental than another. Different qualifications of ‘fundamental’ come into play only in his two other criteria. In this context he specifies that ‘more fundamental’ means ‘being located on a lower level of an abstract/spatial hierarchy’.

  30. 30.

    A modified version of this paper, which is focused more on the differences between part-whole explanations in biology and in physics, was published in the same year (Love and Hüttemann 2011).

  31. 31.

    In fact, I do not see a good reason why Hüttemann and Love should not do that, too. Taking into account the diversity of reductive reasoning and seeking an account of reductive explanation that is sensitive to actual biological practice contradicts neither calling this account a “theory” of reductive explanations nor regarding it as an alternative to other accounts of epistemic reduction.

  32. 32.

    However, in the next section I point out that it remains unclear whether Hüttemann and Love identify reductive explanations with part-whole explanations or whether they assume that there exist also non-reductive part-whole explanations, as some of their remarks suggest.

  33. 33.

    Although I agree with them that part-whole explanations are paradigmatic cases of reductive explanations, in my view it is also important to recognize other features of reductive explanations than their part-whole character as well as other kinds of reductive explanations (see Chap. 6).

  34. 34.

    Hüttemann and Love regard the temporal nature of reductive explanations in biology as one reason why traditional concepts like identity, multiple realization, and supervenience (which are atemporal concepts) are of limited significance to biology (2011, 526, 533).

  35. 35.

    This is important to note because, on the one hand, the temporal character distinguishes biological from physical part-whole explanations and, on the other hand, once part-whole relations are “treated temporally, intrinsicality and fundamentality take on independent significance in reductive explanations” (2011, 530).

  36. 36.

    Hüttemann and Love exclude the possibility that only fundamentality is violated, whereas intrinsicality is still satisfied (and, thus, the explanation can be characterized as reductive). They argue that this is not possible because “if a feature is intrinsic to S then in order to be contained within S it must be instantiated on a more fundamental level than S itself” (2011, fn. 12). As I will object in Sect. 2.3.4, this is true only if one interprets fundamentality according to Sarkar’s “spatial hierarchy” criterion and not if one presupposes Hüttemann’s and Love’s qualified sense of fundamentality (see Sect. 2.3.2).

  37. 37.

    One needs to be careful at this point since Hüttemann and Love claim that there is not just one reductive character of explanations. Rather, different respects in which an explanation can be reductive or not need to be distinguished. Accordingly, they would have done better to have introduced two notions of reductivity, namely reductivityF (for reductivity with respect to fundamentality) and reductivityIF (for reductivity with respect to intrinsicality and fundamentality). The explanation discussed above is only reductiveF, but not reductiveIF.

  38. 38.

    Strictly speaking, they claim that only fundamentality is decoupled from intrinsicality, not vice versa.

  39. 39.

    At least if the notion of a spatial level is defined locally (see Chap. 6).

  40. 40.

    This does not mean that in any part-whole explanation each step of the causal process from t to t* is described. Sometimes there exist “gaps’” in the causal process that is cited as explanatorily relevant. These gaps are due to the unavailability of knowledge or to certain explanatory interests. Nevertheless, even in those cases the explanans is not restricted to the properties of the parts to one particular time t.

  41. 41.

    I suppose that this is why Love and Hüttemann add the bracket “(or t1, …, tn < t*)” to claims such as “the state or properties of the parts and their interactions at t (or t 1 , …, t n < t*) bring about a change in the state or properties of the compound at time t*” (2011, 188) in their second paper on “Comparing Part-Whole Reductive Explanations in Biology and Physics”.

  42. 42.

    The alternative would be to claim that explanations that violate intrinsicality and fundamentality, are neither reductive explanations nor part-whole explanations.

References

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

    Google Scholar 

  • Bechtel, W. (2008). Mental mechanisms. Philosophical perspectives on cognitive neuroscience. New York: Taylor and Francis Group.

    Google Scholar 

  • Craver, C. F. (2002b). Interlevel experiments and multilevel mechanisms in the neuroscience of memory. Philosophy of Science, 69, 83–97.

    Article  Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  • Craver, C. F., & Darden, L. (2005). Introduction. Studies in History and Philosophy of Biological and Biomedical Sciences, 36, 233–244.

    Article  Google Scholar 

  • Dobzhansky, T. (1973). Nothing in biology makes sense except in the light of evolution. American Biology Teacher, 35, 125–129.

    Article  Google Scholar 

  • Dupré, J. (1993). The disorder of things. Metaphysical foundations of the disunity of science. Cambridge: Harvard University Press.

    Google Scholar 

  • Glennan, S. S. (2002). Rethinking mechanistic explanation. Philosophy of Science, 69, 342–353.

    Article  Google Scholar 

  • Hove, J. R., Köste, R. W., Forouhar, A. S., Acevedo-Bolton, G., Fraser, S. E., & Gharib, M. (2003). Intracardiac fluid forces are an essential epigenetic factor for embryonic cardiogenesis. Nature, 421, 172–177.

    Article  Google Scholar 

  • Hull, D. (1974). The philosophy of biological science. Englewood Cliffs: Prentice-Hall.

    Google Scholar 

  • Hüttemann, A., & Love, A. C. (2011). Aspects of reductive explanation in biological science: Intrinsicality, fundamentality, and temporality. British Journal for Philosophy of Science, 62(3), 519–549.

    Article  Google Scholar 

  • Kaiser, M. I. (2012). Why it is time to move beyond nagelian reduction. In D. Dieks, W. J. Gonzalez, S. Hartmann, M. Stöltzner, & M. Weber (Eds.), Probabilities, laws, and structures (The philosophy of science in a European perspective, Vol. 3, pp. 255–272). Heidelberg: Springer.

    Google Scholar 

  • Kaiser, M. I. (forthcoming a). Individuating part-whole relations in the biological world. In O. Bueno, R.-L. Chen, & M. B. Fagan (Eds.), Individuation across experimental and theoretical sciences, Oxford University Press.

    Google Scholar 

  • Kauffmann, S. A. (1970). Articulation of parts explanation in biology and the rational search for them. PSA, 1970, 257–272.

    Google Scholar 

  • Kitcher, P. (1984). 1953 an all that: A tale of two sciences. Philosophical Review, 93, 335–373.

    Article  Google Scholar 

  • Love, A. C. (2008b). Review symposium: More worry and less love. Metascience, 17, 1–26.

    Article  Google Scholar 

  • Love, A. C., & Hüttemann, A. (2011). Comparing part-whole reductive explanations in biology and physics. In D. Dieks, W. J. Gonzalez, S. Hartmann, T. Uebel, & M. Weber (Eds.), Explanation, prediction, and confirmation. New trends and old ones reconsidered (pp. 183–202). Berlin: Springer.

    Chapter  Google Scholar 

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

    Article  Google Scholar 

  • Mayr, E. (1982). The growth of biological thought. Diversity, evolution, and inheritance. Cambridge: Belknap Press.

    Google Scholar 

  • Mitchell, S. D. (2009). Unsimple truths. Science, complexity, and policy. Chicago: University of Chicago Press.

    Book  Google Scholar 

  • Nagel, E. (1961). The structure of science. Problems in the logic of scientific explanation. London: Routledge.

    Google Scholar 

  • Railton, P. (1981). Probability, explanation, and information. Synthese, 48, 233–256.

    Article  Google Scholar 

  • Rosenberg, A. (1985). The structure of biological science. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Rosenberg, A. (1994). Instrumental biology or the disunity of science. Chicago: University of Chicago Press.

    Google Scholar 

  • Rosenberg, A. (2006). Darwinian reductionism. Or, how to stop worrying and love molecular biology. Cambridge: University of Chicago Press.

    Book  Google Scholar 

  • Salmon, W. C. (1993). Scientific explanation and the causal structure of the world. In D. Hillel-Ruben (Ed.), Explanation. Oxford readings in philosophy (pp. 78–112). Oxford: Oxford University Press.

    Google Scholar 

  • Sarkar, S. (1992). Models of reduction and categories of reductionism. Synthese, 91, 167–194.

    Article  Google Scholar 

  • Sarkar, S. (1998). Genetics and reductionism. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Sarkar, S. (2005). Molecular models of life. Philosophical papers on molecular biology. Cambridge: MIT Press.

    Google Scholar 

  • Sarkar, S. (2008). Reduction. In S. Psillos & M. Curd (Eds.), The Routledge companion to philosophy of science (pp. 425–434). London: Routledge.

    Google Scholar 

  • Sarkar, S. (2009). Ecology. In E. N. Zalta (Ed.), Stanford encyclopedia of philosophy (Spring 2009 Edition). http://plato.stanford.edu/archives/spr2009/entries/ecology/

  • Sober, E. (1999). The multiple realizability argument against reductionism. Philosophy of Science, 66, 542–564.

    Article  Google Scholar 

  • Weber, M. (2008). Critical notice: Darwinian reductionism. Biology and Philosophy, 23(1), 143–152.

    Article  Google Scholar 

  • Wimsatt, W. C. (1976a). Reductive explanation: A functional account. PSA, 1974, 671–710.

    Google Scholar 

  • Wimsatt, W. C. (2006a). Reductionism and its heuristics: Making methodological reductionism honest. Synthese, 151(3), 445–475.

    Article  Google Scholar 

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

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Kaiser, M.I. (2015). Two Perspectives on Explanatory Reduction. In: Reductive Explanation in the Biological Sciences. History, Philosophy and Theory of the Life Sciences. Springer, Cham. https://doi.org/10.1007/978-3-319-25310-7_4

Download citation

Publish with us

Policies and ethics