Skip to main content
Log in

Feyerabend’s well-ordered science: how an anarchist distributes funds

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

To anyone vaguely aware of Feyerabend, the title of this paper would appear as an oxymoron. For Feyerabend, it is often thought, science is an anarchic practice with no discernible structure. Against this trend, I elaborate the groundwork that Feyerabend has provided for the beginnings of an approach to organizing scientific research. Specifically, I argue that Feyerabend’s pluralism, once suitably modified, provides a plausible account of how to organize science. These modifications come from C.S. Peirce’s account of the economics of theory pursuit, which has since been corroborated by empirical findings in the social sciences. I go on to contrast this approach with the conception of a ‘well-ordered science’ as outlined by Kitcher (Science, truth, and democracy, Oxford University Press, New York, 2001), Cartwright (Philos Sci 73(5):981–990, 2006), which rests on the assumption that we can predict the content of future research. I show how Feyerabend has already given us reasons to think that this model is much more limited than it is usually understood. I conclude by showing how models of resource allocation, specifically those of Kitcher (J Philos 87:5–22, 1990), Strevens (J Philos 100(2):55–79, 2003) and Weisberg and Muldoon (Philos Sci 76(2):225–252, 2009), unwittingly make use of this problematic assumption. I conclude by outlining a proposed model of resource allocation where funding is determined by lottery and briefly examining the extent to which it is compatible with the position defended in this paper.

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

(taken from Weisberg & Muldoon 2009, 230)

Similar content being viewed by others

Notes

  1. Evolutionary accounts of science suppose that the order is maintained by selective pressures. Some, such as Popper (1972), suppose that these pressures come from methodology; explanatory realists presume that order is maintained by the impingements of the external world; Kuhnians suppose that pressures come from institutions (cf. Wray 2015). Regardless of the source of the pressures, all of these positions maintain that science is ‘ordered.’ See Allchin (2015) for a general discussion of ‘naturalistic’ approaches to organizing science via the ‘self-correcting hypothesis’.

  2. See Barseghyan (2015, p. 35) and Barseghyan and Shaw (2017) for descriptions of what is meant by pursuit.

  3. Kitcher restricts his ideal to public research stating, quite crudely, “privatization of scientific research will probably make matters worse” (Kitcher 2001, p. 126). See Pinto (2015) for a criticism of this.

  4. This view is adopted and slightly modified by Cartwright (2006).

  5. This point has been realized before, and discussed extensively by Poincaré (1902). Poincaré solves the problem of excessive facts with a distinction between simple facts, which have greater epistemic capabilities, and ordinary facts whereas Kitcher appeals to values to determine the significance of facts.

  6. This is explicitly a normative claim and isn’t descriptive of all scientific practices. See Feyerabend (1965a, pp. 156–157) for an overview of the museum set up by the Royal Society that displayed random facts without any sort of ‘selective principle.’ This included the silliest ‘facts’ imaginable, such as the following: “1661, July 24: a circle was made with powder of unicorn’s horn, and a spider set in the middle of it, but it immediately ran out several times repeated. This spider once made some stay upon the powder” (Weld 1848/2011, p. 219). See pg. 93 for an appraisal of the museum as being “necessary for the welfare of science”.

  7. This damnation of A is unfair since any metaphysics, including atheistic metaphysics, will provide criteria for what constitute ‘fundamental truths’ (or lack thereof). Even Kitcher’s earlier view of realism (cf. chap. 5 of Kitcher (1993)) commits him to the view since the best theories will provide us with sets of epistemic relations from which we can discern significance. A stronger justification comes from Kitcher’s adoption of pragmatism (cf. Kitcher 2013) where he accepts William James’ claim that we evaluate methods based on their consequences rather than their foundations (see James 1907 and Dewey 1908 for a critical discussion).

  8. To be clear, throughout this paper I assume that any funding distribution will be two-tiered in that it involves both the salaries of scientists and the external funding of their projects. While most are not explicit about this, discussions of resource allocation are restricted to the latter.

  9. Kitcher, at other points (cf. Kitcher 1997, pp. 296–267), argues that should not pursue lines of research that would interfere with the prospects of underprivileged groups pursuing their conception of the good life (e.g., sociobiology). I think these constraints have already been adequately responded to by Longino (2002) and Eigi (2012).

  10. Toulmin, at times, adopts this view as well (Toulmin 1976, p. 661).

  11. There are some, such as Zahar (1982, p. 406), who read Lakatos as defending a Polanyiian ‘tacit knowledge’ methodology in his later career, despite Lakatos’ fierce criticisms of Polanyi’s ‘elitism’ (cf. Lakatos 1978b, pp. 111–117).

  12. Kitcher repeats this view later in his career (Kitcher 2004). I imagine examples like those of Douglas (2000), where the minutia of scientific reasoning are susceptible to value judgments, would be included under the broad category of what kinds of science should be done.

  13. See Epstein’s (1995) discussion of the political debates over the construction of AIDS trials in California during the 1960 s and 70 s for an example of this.

  14. The same point is made in Flory and Kitcher (2004). Their focus on medical research, however, should be seen as a specific instance of a broader category of scientific research that addresses the morally optimal needs.

  15. See Barseghyan (2015, pp. 30–42) for a discussion of this.

  16. I also take it that methods have content, indirectly at least, since they rely on theoretical assumptions.

  17. Another model that has garnered increasing interest is Zollman’s which is based largely on decision theory and does not, to my knowledge, make use of the projection postulate (Zollman 2010). However, he does claim that “[e]xtending this model to cases where the probability of future success is determined by past success will be left to future research” (Zollman 2010, p. 24). If I am correct, this line of research may be a dead end.

  18. See Frankel (1979) for a discussion of the early development of continental drift.

  19. This is a non-negligible feature of scientific discovery. The locus classicus discussion of serendipity in discovery can be found in Roberts (1989) which has been increasingly confirmed in more recent investigations (García 2009; Hargrave-Thomas et al. 2012). For a more skeptical discussion, see (Jeste et al. 1979).

  20. Topics can defined broadly or narrowly. The WM model is aimed to represent “the topic that a specialized research conference or advanced level monograph might be devoted to” (Weisberg and Muldoon 2009, p. 228).

  21. Two assumptions have already been criticized. Thoma (2015) criticizes the assumption that agents can only move locally (i.e., coding of the controls have “those agents behaving like lethargic random walkers” (433) within their Moore neighborhood), and that the purported benefits of the division of labor are more limited in higher dimensional landscapes 425) especially given the fine-graininess of the distinction between adjacent patches (462). Alexander et al. (2015) criticize the breadth of kinds of agents chosen. These criticisms are distinct from the ones I am concerned with in this paper.

  22. It is possible that Feyerabend’s aversion to the projection postulate is grounded in Popper’s arguments against historicism where we “cannot predict, by rational or scientific methods, the future growth of our scientific knowledge” (Popper 1957, p. vi). This influence is only speculative though, since Feyerabend never, to my knowledge, acknowledges his indebtedness to this view or endorses Popper’s criticisms of historicism more generally. Since Popper’s influence on Feyerabend was mixed, even at the beginning of his career (cf. Collodel 2016), we cannot automatically assume that this argument influenced Feyerabend’s arguments against the projection postulate.

  23. Feyerabend’s most sustained discussion of this phenomena can be found in Chapter 4 of Against Method and is also apparent in his discussion of the revival of classical mircrophysics in the 1960s (Feyerabend 1970a). Additional examples and discussion can be found in Schaffer (1994).

  24. I will speak here of the ‘value’ of ideas for ease of expression though, for obvious reasons, the content of an idea is closely linked to their value.

  25. Additionally, there are many examples of ideas that had intended practical implications and were co-opted to have different practical implications at later points. See De Laet and Mol’s (2000) discussion of the Zimbabwe Bush Pump for a classic example of this.

  26. Notice that if one follows Quine and many realists and argue that all we can ever know about the structure of the world is what we are committed to by our best scientific theories, then no principle could be invoked here since it would require knowledge of the content of future theories.

  27. Feyerabend himself defends the views that our metaphysical theories should contradict, rather than be consistent with, scientific theories (Feyerabend 1965a, p. 183). However, this argument is meant to defend the use of metaphysics for future discoveries as a kind of heuristic and not because we think those theories are actually true.

  28. This argument, pace some commentators (Farrell 2003, pp. 157, 210), extends to contextual rules of reasoning as well (cf. Feyerabend 1977, fn. 1 368).

  29. Popper never extended the principle of testability to his own falsificationism. Lakatos claims that Popper has given an implicit answer, but he provides weak textual evidence to support this interpretation (Lakatos 1970). Feyerabend does, and so does Bartley (1962). See Watkins (1971) for a critical discussion of Bartley’s position.

  30. See Longino (1990, Chapters 1–3) for a discussion of proceduralist methodologies versus ‘justificationist’ approaches to methodology.

  31. Feyerabend actually provides a second, slightly different, argument in the following chapters of Against Method where he shows that counterinduction plays a certain role that induction cannot play (cf. Feyerabend 1975, Chapters 3–5). The claim that methodological pluralism is necessary for maximal testability and the claim that methodological pluralism is necessary because different methods have different functions are distinct.

  32. See Stanford (2006, p. 174) for additional examples of scientists whose judgments were, from a retrospective point of view, mistaken.

  33. Similar results were found in ethnographic studies of the European Research Council (Luukkonen 2012).

  34. I should mention that there are attempts to improve peer-review. For instance, Jayasinghe et al. (2003) have found various markers for referee bias (North American reviewers give higher reviews than Australian reviewers, reviewers nominated by the researcher give higher reviews that those nominated by a granting agency, and scientists with fewer proposals tend to give higher ratings). However, these results still only improve reliability measures to 0.47 (the standard threshold for reliability is normally in between 0.8-0.9 (Marsh et al. 2008, 162)). This being said, enough tenacity on this research program may allow us to reformulate and better implement peer-review.

  35. Roughly, Kuhn provides three primary reasons why this is the case. First, scientists are “freed” from the need to constantly renegotiate their fundamental convictions and focus exclusively on more esoteric phenomena (Kuhn 1962, p. 163). They do not need to worry about justifying the importance of their problem as they would have had they faced a heterodox audience. Secondly, progress is more assured since the past successes of the paradigm give the experimenter some assurance that there is a solution to their problem, a guarantee that isn’t available when experimenting within new frameworks.3 Finally, Kuhn lays out pedagogical reasons why paradigms are more conducive to progress. Rather than have students introduced to their fields by reading their respective “classics” (e.g. Newton’s Principia, Darwin’s Origin of Species, etc.), their textbooks already presuppose these accomplishments in formulating more precise and obscure questions. As such, graduate students are able to contribute to the field more quickly than if they had to learn the theoretical and philosophical foundations of their field (164). See also Kuhn (1959, 1963) for more extended discussions.

  36. While Feyerabend often appears to ground this claim in its intuitive force, he also cites a variety of empirical studies to support this view. It is extremely difficult to assess how this claim fares in light of more contemporary empirical findings, but some preliminary discussions on this topic suggest that the psychological benefits of theoretical diversity Feyerabend promises are both reliable and robust (Preston 2005).

  37. Such admissions can be found in Feyerabend’s earlier writings as well (cf. Feyerabend 1962, pp. 72–75). Feyerabend did not just rely on intuitions and anecdotes, but argued that these theses were empirically supported as well, though they also had some minor drawbacks (cf. Feyerabend 1965b, 130–132, 1970b, fn. 42 107).

  38. The quotes within this quote come from Mill, whom Feyerabend leaned on heavily in his later career. For a discussion of Feyerabend’s relationship to Mill, see Lloyd (1997).

  39. For a critical discussion of this argument, see Preston (1997, pp. 35–36) for a discussion and Zahar (1982).

  40. Feyerabend repeats this principle verbatim in the paper version of Against Method (Feyerabend 1970b, p. 25).

  41. Feyerabend’s definition of ‘theory’ is explicitly broad enough to include Kuhn’s notion of ‘paradigm’.

  42. Additionally, Feyerabend argues that the significance of refuting instances of theories change as the result of the presence of an alternative explanation (cf. Feyerabend 1965b, fn. 7 106).

  43. Feyerabend first became aware of this issue in his correspondence with Kuhn (see the letters published in Hoyningen-Huene 1995, 2006) in the early 1960s.

  44. Motterlini (1999, pp. 3–4) provides a nice reconstruction of Feyerabend and Lakatos’ dialogue about the possibility of a ‘time limit.’ Ultimately, Feyerabend argues that the original motivation for tenacity, to provide ‘breathing space’ for theories, remains true at any stage in a theories development (cf. Feyerabend 1970d, p. 215).

  45. Some, such as proponents of the strong programme, may intentionally conflate empirical questions about science and the content of science. While Feyerabend never directly addresses this question, his manner of discussing theories suggests that he does not reduce the content of those theories to beliefs of scientists. For a critical discussion of this, see Sismondo (1993) and Brown (2004).

  46. Other interpretations include the ‘generative interpretation’ where abduction provides a means for creating hypothesis (Nickles 1985) and the ‘justificatory interpretation’ where abduction serves as a means to infer general fallible knowledge claims (Misak 2000). I will not take a stance on which interpretation most accurately reflects Peirce’s ‘real’ views here.

  47. See Radnitzky (1987) for an explication of falsificationism in economic terms.

  48. Kuhn (1963) recognizes and stresses this point. However, he argues that the paradigm in place is exclusive in that proliferation is only to be engaged in during periods of revolution.

  49. Before detailing these observations, it should be noted that there are also serious limitations with these models such as the inability to forecast future scientific labor markets (Leslie and Oaxaca 1993). Most estimated cost models also presuppose that technological costs will decrease linearly making substantive assumptions about progress in technology that aren’t well studied empirically.

  50. There are also several examples of cheap experiments which had major initial impacts which were then followed up with more sophisticated, and expensive, experiments. For example, V.S. Ramachandran’s early experiments on phantom limbs were done with cardboard boxes and ordinary mirrors that have now become refined in virtual reality experiments. Scherer (1966) suggests that this trend is fairly robust and is also hinted at in Peirce’s example of Wollaston.

  51. It is unclear, to me at least, whether ‘duplications’ are the same as ‘replications.’ If they are, fields undergoing a ‘replication crisis’ may consider this to be a moot point.

  52. See Stephan (1996, pp. 1216–1217) and the citations therein.

  53. See Feyerabend (1964) for his criticisms of induction.

  54. “If they contradict a well-confirmed point of view, then this indicates their usefulness as an alternative. Alternatives are needed for the purpose of criticism. Hence metaphysical systems that contradict observational results or well-confirmed theories are most welcome starting points of such criticism. Far from being misfired attempts at anticipating, or circumventing, empirical research that have been deservedly exposed by reference to experiment, they are the only means we possess for examining the assumptions implicit in our observational results” (Feyerabend 1965a, p. 183).

  55. I have bracketed concerns about justice and egalitarianism from this paper, though such concerns are of the utmost importance.

  56. Herbert et al. (2013) estimate that 550 working years went into proposals submitted to the NHMRC in March, 2012. This is on top of the opportunity cost of reviewer hours and the administrative costs of funding bodies.

  57. I am willing to admit that there is some grey area around the line of the cranks. After all, some reviewers are harsher than others. I don’t think this grey area extends very far and, therefore, we can simplify out discussion by discussing a strict dichotomy between the cranks and reasonable research.

  58. Though see Feyerabend (2011, pp. 45–46) for some qualifications.

  59. In the case of the Human Brain Project, many non-computational approaches in the mind-brain sciences lost funding and it diverted resources to solely pursuing computational approaches (of a very specific style). See Shaw (2018, pp. 89–93) for a discussion of this.

References

  • Achinstein, P. (2000). Proliferation: is it a good thing? In J. Preston, G. Munévar, & D. Lamb (Eds.), The worst enemy of science? Essays in memory of Paul Feyerabend (pp. 37–57). Oxford: Oxford University Press.

    Google Scholar 

  • Alexander, J. M., Himmelreich, J., & Thompson, C. (2015). Epistemic landscapes, optimal search, and the division of cognitive labor. Philosophy of Science, 82(3), 424–453.

    Google Scholar 

  • Allchin, D. (2015). Correcting the ‘self correcting’ mythos of science. Filosofiae Historia da Biologia, Sao Paulo, 10(1), 19–35.

    Google Scholar 

  • Avin, S. (2015a). Breaking the grant cycle: On the rational allocation of public resources to scientific research projects. Doctoral Dissertation, University of Cambridge.

  • Avin, S. (2015b). Funding Science by Lottery. In recent developments in the philosophy of science: EPSA13 Helsinki (pp. 111–126). Berlin: Springer.

    Google Scholar 

  • Barseghyan, H. (2015). The laws of scientific change. New York: Springer.

    Google Scholar 

  • Barseghyan, H., & Shaw, J. (2017). Can a taxonomy of stances clarify classic debates about scientific change? Philosophies, 2(4), 24.

    Google Scholar 

  • Bartley, W. (1962). The retreat to commitment. Berkeley: University of California Press.

    Google Scholar 

  • Brown, W. (1983). The economy of Peirce’s abduction. Transactions of the Charles S. Peirce Society, 19(4), 397–411.

    Google Scholar 

  • Brown, J. (2004). Who rules in science?. Cambridge: Harvard University Press.

    Google Scholar 

  • Cartwright, N. (2006). Well-ordered science: Evidence for use. Philosophy of Science, 73(5), 981–990.

    Google Scholar 

  • Cole, S., Cole, J., & Simon, G. (1981). Chance and consensus in peer review. Science, 214, 20.

    Google Scholar 

  • Collodel, M. (2016). Was Feyerabend a Popperian? Methodological issues in the history of the philosophy of science. Studies in History and Philosophy of Science Part A, 57, 27–56.

    Google Scholar 

  • De Laet, D., & Mol, A. (2000). The Zimbabwe bush pump: Mechanics of a fluid technology. Social Studies of Science, 30(2), 225–263.

    Google Scholar 

  • Dewey, J. (1908). What does pragmatism mean by practical? The Journal of Philosophy, Psychology and Scientific Methods, 5(4), 85–99.

    Google Scholar 

  • Douglas, H. (2000). Inductive risk and values in science. Philosophy of Science, 67(4), 559–579.

    Google Scholar 

  • Edgeworth, F. (1888). The statistics of examinations. Journal of the Royal Statistical Society, 51(3), 599–635.

    Google Scholar 

  • Edgeworth, F. (1890). The element of chance in competitive examinations. Journal of the Royal Statistical Society, 53(3), 460–475.

    Google Scholar 

  • Eigi, J. (2012). Two millian arguments: Using Helen Longino’s approach to solve the problems Philip Kitcher targeted with his argument on freedom of inquiry. Studia Philosophica Estonica, 5(1), 44–63.

    Google Scholar 

  • Epstein, S. (1995). The construction of lay expertise: AIDS activism and the forging of credibility in the reform of clinical trials. Science, Technology and Human Values, 20(4), 408–437.

    Google Scholar 

  • Farrell, R. (2003). Feyerabend and scientific values: Tightrope-walking rationality. Dordrecht: Kluwer.

    Google Scholar 

  • Feyerabend, P. (1958). An attempt at a realistic interpretation of experience. In Proceedings of the aristotelian society (Vol. 58, pp. 143–170). Aristotelian Society, Wiley.

  • Feyerabend, P. (1960). On the interpretation of scientific theories. Logic, theory of knowledge, philosophy of science, philosophy of language. In Proceedings of the 12th International Congress of Philosophy, Venice, 12-18 September 1958, (Vol. 5, pp. 151–159). Florence: Sansoni.

  • Feyerabend, P. (1961). “Comments on Grünbaum’s ‘law and convention in physical theory’”. In current issues in the philosophy of science: symposia of scientists. In H. Feigl & G. Maxwell (Eds.), Current issues in the philosophy of science: Symposia of Scientists and philosophers, proceedings of section L of the American association for the advancement of science, 1959 (pp. 155–161). New York: Holt, Rinehart and Winston.

    Google Scholar 

  • Feyerabend, P. (1962). Explanation, reduction and empiricism. In H. Feigl & G. Maxwell (Eds.), Scientific explanation, space and time, Minnesota studies in the philosophy of science (Vol. III, pp. 28–97). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Feyerabend, P. (1963). How to be a good empiricist: A plea for tolerance in matters epistemological. In B. Baumrin (Ed.), Philosophy of science: The Delaware seminar (Vol. 2, pp. 3–39). New York City: Interscience Publishers.

    Google Scholar 

  • Feyerabend, P. (1964). A note on two ‘problems’ of induction. The British Journal for the Philosophy of Science, 19(3), 251–253.

    Google Scholar 

  • Feyerabend, P. (1965a). Problems of empiricism. In R. Colodny (Ed.), Beyond the edge of certainty: Essays in contemporary science and philosophy. University of Pittsburgh series in the philosophy of science (Vol. 2, pp. 145–260). Englewood Cliffs, NJ: Prentice-Hall.

    Google Scholar 

  • Feyerabend, P. (1965b). Reply to criticism: Comments on smart, sellars and putnam. In Proceedings of the Boston colloquium for the philosophy of science (pp. 223–261).

  • Feyerabend, P. (1966a). Dialectical materialism and the quantum theory. Slavic Review, 25(3), 414–417.

    Google Scholar 

  • Feyerabend, P. (1966b). On the possibility of a perpetuum mobile of the second kind. In P. Feyerabend & G. Maxwell (Eds.), Mind, matter and method: Essays in philosophy and science in Honor of Herbert Feigl (pp. 3–13). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Feyerabend, P. (1968). Outline of a pluralistic theory of knowledge and action. In S. Anderson (Ed.), Planning for diversity and choice (pp. 275–284). Cambridge: MIT Press.

    Google Scholar 

  • Feyerabend, P. (1970a). In defence of classical physics. Studies in History and Philosophy of Science, 1, 59–85.

    Google Scholar 

  • Feyerabend, P. (1970b). Against method: Outline of an anarchistic theory of knowledge. In M. Radner & S. Winokur (Eds.), Analysis of theories and methods of physics and psychology, Minnesota studies in the philosophy of science (Vol. 4, pp. 17–130). Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Feyerabend, P. (1970c). Classical empiricism. In R. E. Butts & J. W. Davis (Eds.), The methodological heritage of newton (Vol. 4, pp. 150–170). Oxford: Basil Blackwell.

    Google Scholar 

  • Feyerabend, P. (1970d). Consolations for the specialist. In I. Lakatos & A. Musgrave (Eds.), Criticism and the growth of knowledge (pp. 197–231). Cambridge: Cambridge University Press.

    Google Scholar 

  • Feyerabend, P. (1975). Against method (1st ed.). London: Verso Books.

    Google Scholar 

  • Feyerabend, P. (1976). On the critique of scientific reason. In R. S. Cohen, P. K. Feyerabend, & M. W. Wartofsky (Eds.), Essays in memory of Imre Lakatos (pp. 109–143). Boston: Reidel.

    Google Scholar 

  • Feyerabend, P. (1977). Changing patterns of reconstruction. The British Journal for the Philosophy of Science, 28(3), 351–369.

    Google Scholar 

  • Feyerabend, P. (1981a). Proliferation and realism as methodological principles. In rationalism, realism, and scientific method: Philosophical papers (Vol. 1, pp. 139–145). Cambridge: Cambridge University Press.

    Google Scholar 

  • Feyerabend, P. (1981b). Two models of epistemic change. In P. K. Feyerabend (Ed.), Problems of empiricism, philosophical papers (Vol. 2, pp. 65–80). Cambridge: Cambridge University Press.

    Google Scholar 

  • Feyerabend, P. (1988). Against method (3rd ed.). London: Verso Books.

    Google Scholar 

  • Feyerabend, P. (2011). The tyranny of science. Cambridge: Polity Press.

    Google Scholar 

  • Flory, J., & Kitcher, P. (2004). Global health and the scientific research agenda. Philosophy & Public Affairs, 32(1), 36–65.

    Google Scholar 

  • Frankel, H. (1979). The career of continental drift theory: An application of Imre Lakatos’ analysis of scientific growth to the rise of drift theory. Studies in History and Philosophy of Science Part A, 10(1), 21–66.

    Google Scholar 

  • García, P. (2009). Discovery by serendipity: A new context for an old riddle. Foundations of Chemistry, 11(1), 33–42.

    Google Scholar 

  • Gillies, D. (2008). How should research be organized?. London: College Publications.

    Google Scholar 

  • Gillies, D. (2014). Selecting applications for funding: Why random choice is better than peer review. A Journal on Research Policy and Evaluation, 2, (1).

    Google Scholar 

  • Goodwin, B. (2005). Justice by lottery (2nd ed.). Charlottesville: Imprint Academic.

    Google Scholar 

  • Graves, N., Barnett, A., & Clarke, P. (2011). Funding grant proposals for scientific research: Retrospective analysis of scores by members of grant review panel. British Medical Journal, 343, d4797.

    Google Scholar 

  • Hargens, L., & Herting, J. (1990). Neglected considerations in the analysis of agreement among journal referees. Scientometrics, 19(1–2), 91–106.

    Google Scholar 

  • Hargrave-Thomas, E., Yu, B., & Reynisson, J. (2012). Serendipity in anticancer drug discovery. World Journal of Clinical Oncology, 3(1), 1.

    Google Scholar 

  • Herbert, D. L., Barnett, A. G., Clarke, P., & Graves, N. (2013). On the time spent preparing grant proposals: An observational study of australian researchers. British Medical Journal, 3(5), e002800.

    Google Scholar 

  • Hoyningen-Huene, P. (1995). Two letters of Paul Feyerabend to Thomas S. Kuhn on a draft of the structure of scientific revolutions. Studies in History and Philosophy of Science Part A, 26(3), 353–387.

    Google Scholar 

  • Hoyningen-Huene, P. (2006). More letters by Paul Feyerabend to Thomas S. Kuhn on proto-structure. Studies in History and Philosophy of Science Part A, 37(4), 610–632.

    Google Scholar 

  • James, W. (1907). Pragmatism, a new name for some old ways of thinking, popular lectures. Green: Longmans.

    Google Scholar 

  • Jayasinghe, U., Marsh, H., & Bond, N. (2003). A multilevel cross-classified modelling approach to peer review of grant proposals: The effects of assessor and researcher attributes on assessor ratings. Journal of the Royal Statistical Society: Series A (Statistics in Society), 166(3), 279–300.

    Google Scholar 

  • Jeste, D., Gillin, J., & Wyatt, R. (1979). Serendipity in biological psychiatry: A myth? Archives of General Psychiatry, 36(11), 1173–1178.

    Google Scholar 

  • Kitcher, P. (1990). The division of cognitive labor. The Journal of Philosophy, 87, 5–22.

    Google Scholar 

  • Kitcher, P. (1993). The advancement of science. New York City: Oxford University Press.

    Google Scholar 

  • Kitcher, P. (1997). An argument about free inquiry. Noûs, 31(3), 279–306.

    Google Scholar 

  • Kitcher, P. (2001). Science, truth, and democracy. New York City: Oxford University Press.

    Google Scholar 

  • Kitcher, P. (2004). What kinds of science should be done? In A. Lightman, D. Sarewitz, & C. Desser (Eds.), Living with the genie (pp. 201–224). Washington, D.C.: Island Press.

    Google Scholar 

  • Kitcher, P. (2013). Toward a pragmatist philosophy of science. THEORIA. Revista de Teoría, Historia y Fundamentos de la Ciencia, 28(2), 185–231.

    Google Scholar 

  • Kuhn, T. (1959). The essential tension: Tradition and innovation in scientific research? The Essential Tension (pp. 225–240). Chicago: University of Chicago Press.

    Google Scholar 

  • Kuhn, T. (1962). The structure of scientific revolutions. Chicago: The University of Chicago Press.

    Google Scholar 

  • Kuhn, T. (1963). The function of dogma in scientific research. In Scientific change: Proceedings of the symposium on the history of science (pp. 347–369).

  • Lakatos, I. (1970). Falsification and the methodology of scientific research programmes. In I. Lakatos & A. Musgrave (Eds.), Criticism and the growth of knowledge (pp. 91–197). Cambridge: Cambridge University Press.

    Google Scholar 

  • Lakatos, I. (1978a). The social responsibility of science. In J. Worrall & G. Currie (Eds.), Philosophical papers, volume 2: Mathematics, science and epistemology (pp. 256–259). Camrbidge: Cambridge University Press.

    Google Scholar 

  • Lakatos, I. (1978b). The problem of appraising scientific theories: Three approaches. In J. Worrall & G. Currie (Eds.), Philosophical papers, volume 2: Mathematics, science and epistemology (pp. 107–121). Camrbidge: Cambridge University Press.

    Google Scholar 

  • Lee, C. (2015). Commensuration bias in peer review. Philosophy of Science, 82(5), 1272–1283.

    Google Scholar 

  • Leslie, L., & Oaxaca, R. (1993). Scientist and engineer supply and demand. Higher Education: Handbook of Theory and Research, 9, 154–211.

    Google Scholar 

  • Lloyd, E. (1997). Feyerabend, mill, and pluralism. Philosophy of Science, 64, S396–S407.

    Google Scholar 

  • Longino, H. (1990). Science as social knowledge: values and objectivity in scientific inquiry. Princeton: Princeton University Press.

    Google Scholar 

  • Longino, H. (2002). Science and the common good: Thoughts on Philip Kitcher’s science, truth, and democracy. Philosophy of Science, 69(4), 560–568.

    Google Scholar 

  • Luukkonen, T. (2012). Conservatism and risk-taking in peer review: Emerging ERC practices. Research Evaluation, 21, 48–60.

    Google Scholar 

  • Marsh, H., Jayasinghe, U., & Bond, N. (2008). Improving the peer-review process for grant applications: Reliability, validity, bias, and generalizability. American Psychologist, 63(3), 160.

    Google Scholar 

  • McKaughan, D. (2008). From ugly duckling to swan: C.S. Peirce, abduction, and the pursuit of scientific theories. Transactions of the Charles S. Peirce Society: A Quarterly Journal in American Philosophy, 44(3), 446–468.

    Google Scholar 

  • Misak, C. (2000). Peirce. In W. Newton-Smith (Ed.), A companion to the philosophy of science (pp. 335–339). Oxford: Blackwell Publishers.

    Google Scholar 

  • Motterlini, M. (Ed.). (1999). For and against method, including Lakatos’s lectures on scientific method, and the Lakatos-Feyerabend correspondence. Chicago: University of Chicago Press.

    Google Scholar 

  • Nickles, T. (1985). Beyond divorce: current status of the discovery debate. Philosophy of Science, 52(2), 177–206.

    Google Scholar 

  • Peirce, C. S. (1879/1967). Note on the theory of the economy of research. Operations Research, 15(4), 643–648.

  • Peirce, C.S. (1958). In: Hartshorne, C., Weiss, P., & Burks, A. (eds), Collected papers of Charles Sanders Peirce (Vol. 8). Cambridge: Harvard University Press. (Massachusetts, 1931–1958; vols. 1–6 edited by C. Harteshorne and Paul Weiss, 1931–1935; vols. 7–8 edited by A. Burks).

  • Peirce, C. S., & Eisele, C. (Eds.). (1976). The new elements of mathematics. The Hague: Mouton.

    Google Scholar 

  • Pinto, M. (2015). Commercialization and the limits of well-ordered science. Perspectives on Science, 23(2), 173–191.

    Google Scholar 

  • Poincaré, H. (1902). Relations between experimental physics and mathematical physics. The Monist, 12, 516–543.

    Google Scholar 

  • Polanyi, M. (1951). The logic of liberty. London: Routledge & Kegan Paul.

    Google Scholar 

  • Popper, K. (1935). The logic of scientific discovery. New York City: Routledge.

    Google Scholar 

  • Popper, K. (1957). The poverty of historicism. New York City: Routledge.

    Google Scholar 

  • Popper, K. (1972). Objective knowledge: An evolutionary approach. Oxford: Clarendon Press.

    Google Scholar 

  • Preston, J. (1997). Feyerabend: Philosophy, science and society. Hoboken: Wiley.

    Google Scholar 

  • Preston, C. (2005). Pluralism and naturalism: Why the proliferation of theories is good for the mind. Philosophical Psychology, 18(6), 715–735.

    Google Scholar 

  • Radnitzky, G. (1987). The ‘economic’ approach to the philosophy of science. British Journal for the Philosophy of Science, 38, 159–179.

    Google Scholar 

  • Reiss, J., & Kitcher, P. (2009). Biomedical research, neglected diseases, and well-ordered science. THEORIA. Revista de Teoría, Historia y Fundamentos de la Ciencia, 24(3), 263–282.

    Google Scholar 

  • Rescher, N. (1978). Peirce and the economy of research. Philosophy of Science, 43(1), 71–98.

    Google Scholar 

  • Roberts, R. (1989). Serendipity: Accidental discoveries in science. Hoboken: Wiley.

    Google Scholar 

  • Schaffer, S. (1994). Making up discovery. In M. A. Boden (Ed.), Dimensions of creativity (pp. 13–51). Cambridge: The MIT Press.

    Google Scholar 

  • Scherer, F. (1966). Time-cost tradeoffs in uncertain empirical research projects. Naval Research Logistics Quarterly, 13(1), 71–82.

    Google Scholar 

  • Shaw, J. (2017). Was Feyerabend an anarchist? The structure(s) of ‘anything goes’. Studies in History and Philosophy of Science Part A, 64, 11.

    Google Scholar 

  • Shaw, J. (2018). Why the realism debate matters for science policy: The case of the human brain project. Spontaneous Generations: A Journal for the History and Philosophy of Science, 8, 82.

    Google Scholar 

  • Sismondo, S. (1993). Some social constructions. Social Studies of Science, 23(3), 515–553.

    Google Scholar 

  • Stanford, K. (2006). Exceeding our grasp: Science, history, and the problem of unconceived alternatives. Oxford: Oxford University Press.

    Google Scholar 

  • Stanford, K. (2015). Unconceived alternatives and conservatism in science: The impact of professionalization, peer-review, and big science. Synthese, 1, 1–18.

    Google Scholar 

  • Stephan, P. (1996). The economics of science. Journal of Economic Literature, 34(3), 1199–1235.

    Google Scholar 

  • Strevens, M. (2003). The role of the priority rule in science. The Journal of Philosophy, 100(2), 55–79.

    Google Scholar 

  • Thoma, J. (2015). The epistemic division of labor revisited. Philosophy of Science, 82(3), 454–472.

    Google Scholar 

  • Toulmin, S. (1976). History, praxis and the ‘third world’: Ambiguities in Lakatos’ theory of methodology. In R. S. Cohen, P. K. Feyerabend, & M. W. Wartofsky (Eds.), Essays in Memory of Imre Lakatos (pp. 655–675). Reidel: Dordrecht and Boston.

    Google Scholar 

  • Watkins, J. (1971). CCR: A refutation. Philosophy, 47, 56–61.

    Google Scholar 

  • Weisberg, M., & Muldoon, R. (2009). Epistemic landscapes and the division of cognitive labor. Philosophy of Science, 76(2), 225–252.

    Google Scholar 

  • Weld, C. R. (1848/2011). A history of the royal society: With memoirs of the presidents (Vol. 1). Cambridge: Cambridge University Press.

  • Wray, K. B. (2015). Kuhn’s social epistemology and the sociology of science. In Kuhn’s structure of scientific revolutions-50 years on (pp. 167–183). Cham: Springer.

  • Zahar, E. (1982). Feyerabend on observation and empirical content. The British Journal for the Philosophy of Science, 33(4), 397–409.

    Google Scholar 

  • Zollman, K. (2010). The epistemic benefit of transient diversity. Erkenntnis, 72(1), 17–35.

    Google Scholar 

Download references

Acknowledgements

Special thanks to Kathleen Okruhlik, Chris Smeenk, Gillian Barker, Anjan Chakravarrty, Eric Desjardins, and two anonymous referees for their helpful feedback on earlier versions of this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jamie Shaw.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shaw, J. Feyerabend’s well-ordered science: how an anarchist distributes funds. Synthese 198, 419–449 (2021). https://doi.org/10.1007/s11229-018-02026-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-018-02026-3

Keywords

Navigation