Skip to main content
Log in

Active externalism, virtue reliabilism and scientific knowledge

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

Combining active externalism in the form of the extended and distributed cognition hypotheses with virtue reliabilism can provide the long sought after link between mainstream epistemology and philosophy of science. Specifically, by reading virtue reliabilism along the lines suggested by the hypothesis of extended cognition, we can account for scientific knowledge produced on the basis of both hardware and software scientific artifacts (i.e., scientific instruments and theories). Additionally, by bringing the distributed cognition hypothesis within the picture, we can introduce the notion of epistemic group agents, in order to further account for collective knowledge produced on the basis of scientific research teams.

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

Similar content being viewed by others

Notes

  1. Apart from some very few exceptions pointing towards the opposite direction (Goldman 2004, 2010; Fuller 2007, 2012), for the most part, even the emerging field of social epistemology has so far focused on the individual epistemic subject, by studying only how individualistic knowledge is affected by social factors. For a general discussion of the field of social epistemology, the debate over its methodology, and its future see (Palermos and Pritchard 2013). For an extended discussion on the compatibility of contemporary approaches within cognitive science with contemporary accounts of knowledge and justification within mainstream and social epistemology, see (Carter et al. 2014). Non-mainstream approaches to social epistemology (e.g., Science and Technology Studies and the Sociology of Scientific Knowledge (Barnes et al. 1996; Bloor 1991; Latour 1999, 2007; Latour and Woolgar 1986) can, no doubt, also prove relevant to bridging the gap between epistemology and philosophy of science and cognitive science. In what, follows, however, we will need to bracket such approaches in order to save space for the discussion of mainstream epistemology, which is the primary focus of the present paper.

  2. See Bach-y-Rita and Kercel (2003) for a recent review on TVSS.

  3. See also (Shani 2013), whose view—viz., moderate active externalism—is similar to what we here call the hypothesis of extended cognition (though note that Shani’s arguments do not so heavily rely on DST, and his view is stronger than the hypothesis of extended cognition in that it denies—instead of remaining silent on the matter—the extension of (common-sense functionalist) mental states. For more details on why common-sense functionalism is necessary for the extended mind thesis, but not the extended and distributed cognition hypotheses, see Palermos (2014a). Again though, note that the hypotheses of extended and distributed cognition are neither incompatible with common-sense functionalism, nor anti-functionalist on the whole. In so far as a cognitive process is a function, these two hypotheses are compatible with functionalism.

  4. Note that the argument of this approach moves from the existence of extended/distributed cognitive processes to the existence of extended/distributed cognitive systems. There are some alternative approaches to active externalism available in the literature, however, that seem to focus either on the relevant extended systems (e.g., (van Gelder 1995; Haugeland 1993, 2000) or on the relevant extended processes (Menary 2013), alone.

  5. To preempt a possible worry, here, the relevant reciprocal interactions need only be continuous during the operation of the relevant coupled cognitive system and the unfolding of any processes related to it. For example, if, as part of her job and during normal working hours, individual \(S\) participates in distributed cognitive system \(X\), \(S\) does not need to continuously interact with the other members of \(X\), when she is at home. However, whenever \(X\) is in operation, \(S\) must continuously and reciprocally interact with the rest of the \(X\)-members. For more details on how dynamical systems theory, in general, can help us clearly distinguish between the hypothesis of extended cognition and the hypothesis of embedded cognition as well as avoid several other worries with respect to the hypothesis of extended cognition (e.g., the ‘cognitive bloat’ worry and the ‘causal-constitution’ fallacy), see Palermos (2014a).

  6. There are several other proponents of virtue reliabilism—most famously Sosa (1988, 1993, 2007). The reason why only the above references have been included in the main text is to indicate a specific lineage of virtue reliabilism that is particularly apt for our present purposes. In the beginning of the line, however, is Greco, who has, himself, been heavily influenced by Sosa’s alternative.

  7. It should be noted that the notion of one’s ‘cognitive character’ as employed by virtue reliabilists is a technical notion, whose meaning may seem counterintuitive to our common-sense understanding of one’s character as consisting of general traits, such as open-mindedness, wisdom, courage, honesty, understanding, empathy and the like. In fact, within mainstream epistemology, there are two main ways in which one can understand intellectual virtues and character: (1) in terms of general ‘trait-virtues’, such as open-mindedness, honesty, understanding and so on and (2) in terms of specialized ‘faculty-virtues’, such as memory, perception, reasoning and so on. Zagzebski (1996) is a typical example of the former approach and Greco (1999, 2010) of the latter. For the distinction between these two approaches to intellectual virtues and thereby character, see Baehr (2006) and Greco and Zagzebski (2000). Since it is the second approach to virtues—in terms of faculties and cognitive abilities—that informs the virtue reliabilist’s sense of one’s ‘cognitive character’, in what follows, the agent’s ‘cognitive character’ refers to the agent’s set of cognitive faculties (as opposed to general, ‘trait-virtues’).

  8. This is a weak formulation of virtue reliabilism for two reasons. First, because it is only a necessary condition on knowledge (several epistemologists hold that virtue reliabilism is a necessary component, but to have an adequate theory of knowledge, they argue, it must be further supplemented by either the safety or the sensitivity principle (Pritchard 2010a)). Second, because, in order to also accommodate testimonial knowledge (Pritchard 2010b), it requires that one’s cognitive success be significantly, as opposed to primarily, creditable to one’s cognitive agency. Accordingly, ‘\(\hbox {COGA}_{weak}\)’ stands for ‘weak COGnitive Agency’ to indicate that this is an account of knowledge, which requires that one’s cognitive success be creditable to one’s cognitive agency only to a significant (as opposed to primary) degree. For more details, see Pritchard (2010b).

  9. For a detailed overview of the epistemic internalism/externalism debate and how it maps onto the internalism/ externalism debate within philosophy of mind see (Carter et al. 2014). For a detailed analysis of the above minimal yet epistemically adequate notion of subjective justification/epistemic responsibility and its relation to cognitive integration see Palermos (2014b).

  10. Elsewhere Palermos (2011, 2014b), it has been argued that both theories put also forward the same broad, common sense functionalist intuitions on what is required from a process to count as a cognitive ability. Briefly, both views state that the process must be (a) normal and reliable, (b) one of the agent’s habits/dispositions and (c) integrated into the rest of the agent’s cognitive character/system.

  11. Making observations through a telescope clearly qualifies as a case of cognitive extension as it is a dynamical process that involves ongoing reciprocal interactions between the agent and the artifact. Moving the telescope around, while adjusting the lenses, generates certain effects (e.g., shapes on the lens of the telescope), whose feedback drives the ongoing cognitive loops along. Eventually, as the process unfolds, the coupled system of the agent and his telescope is able to identify—that is, see—the target satellite.

  12. On the basis of the feedback loops between the agent and his artifact (see the footnote above).

  13. See Berk and Garvin (1984). See also (Bamberger and Brofsky 1979; Olson 1996a, b; Tomasello 2009; Wertsch 1988).

  14. Following Olson (2002, 1996a) [who draws on (Goody 1986, 1987) and (Shankweiler and Liberman 1972)], and contrary to (Saussure 1916/1983) and Bloomfield (1933), we should here note that writing may not be the mere transcription of speech, but a language in its own right. For ideas similar to Olson’s see (Harris 1989) and Linell (2005). For a discussion of how they fit into the current debate over active externalism, see (Theiner 2013) and (Theiner 2011, ch. 4).

  15. In his (2008), Logan claims that our primitive, biological cognitive capacities are all percept-based and that it is only with the advent of language that we acquired concepts. On the face of this it would seem that his approach is a form of concept empiricism. But this may turn out to not stand upon further reflection: First, the debate over empiricism versus rationalism can be construed in several ways ((Markie 2013), see also (Weiskopf 2007) for a recent defense of several weak versions of empiricism, as informed by neuroscientific evidence)); second, and against the background of all these dialectical possibilities, Logan stresses the deeply transformative effects of language on thought, indicating that as thought evolves in response to the transformative effects of language, it may develop into something that is more than just its perceptual origins.

  16. In what follows, I do not draw a clear distinction between theories and their models. The two most prevailing views about the relation between models and theory are (1) the syntactic view (associated with the logical positivists) and (2) the semantic view (Suppes 1961, 1967; Suppe 1977; van Fraassen 1980; Giere 1988). Both of these views understand the relation between models and theory as a particularly close one. According to the syntactic view, models are dictated by theories, whereas on the semantic view, theories just are families of models. Recently however, Morgan and Morrison (1999) have put forward a third alternative according to which models are ‘autonomous agents’ in the sense that they are not entirely reliant either on theories or on the world they are meant to represent. For more details, see (Morgan and Morrison 1999, ch. 2).

  17. As a referee also points out, several authors (e.g.,Goody 1977; Latour and Woolgar 1986; Olson 1996a; Tufte 2001) have in the past claimed that symbolic representations play a crucial role in the development and visualization of scientific knowledge. These authors, however, have not advanced the strong claim that cognition is literally externalized, rather than heavily dependent on external tools and aids, along the lines suggested by the alternative hypothesis of embedded cognition (Adams and Aizawa 2001, 2010; Rupert 2004, 2009). As noted in §2.1, however, the hypothesis of extended cognition should be clearly distinguished from the hypothesis of embedded cognition, which is a conservative view: On the present approach, scientific theories do not act as mere scaffolds of the epistemic agent’s internal cognitive capacities, but they are literally constitutive of the cognitive repertoire that makes up the agent’s cognitive character. Briefly, the reason, according to dynamical systems theory, is that when the dependence between the cognitive agent and the world is one-way, cognition is indeed merely embedded. When a cognitive system interacts reciprocally with some aspect of the world, however, then the two constitute a coupled system that consists of both of them. It is for this reason that, in what follows, the focus will be on cases whereby the epistemic agent mutually interacts in an ongoing way with external scientific symbols.

  18. For the importance of the normative aspects of the external representational systems in explaining cognition see Menary (2007).

  19. An anonymous referee suggests that a similar example with an equally important impact on the development of further scientific theories is the case of Feynman diagrams. Similarly, Kaiser (2005, p. 156) accentuates their role by noting that “Feynman diagrams have revolutionized nearly every aspect of theoretical physics.”

  20. According to Lakatos, science and all scientists proceed on the basis of ‘sophisticated methodological falsificationism’, which is the methodological background against which he presented his rational reconstruction of science.

  21. One possible worry here is to ask whether the above distribution of credit suggests that we should also posit distributed cognitive systems that would be vastly extended not only in space but also back in time. This would be a rather awkward claim to make, but we can clearly answer in the negative: As noted in §2.1, in order to have an extended or distributed cognitive system we need that all the contributing members reciprocally (i.e., non-linearly) interact with each other. The dependency relations that give rise to the distribution of credit I refer to above, however, are all, one-way, linear ones. Accordingly, given the dynamical systems theory approach to extended and distributed cognition we here draw upon, we have no reason to think that the above distribution of credit requires to also posit distributed cognitive systems that may even extend back in time. Thanks to an anonymous referee for requesting me to clarify this potentially confusing point.

  22. Think about Newton’s infamous remark in a letter to his rival, Robert Hooke:

    What Descartes did was a good step. You have added much several ways and especially in taking the colours of thin plates into philosophical consideration. If I have seen further is by standing on the shoulders of Giants.

    Even if, as is often assumed, the above is in fact intended as a sarcastic remark (due to Hooke’s short build), the ambivalence that Newton takes advantage of derives from the truthfulness of the expression within the context it is uttered—that no matter the genius and the individual efforts of a scientist, his or her achievements will always rest upon and derive from the achievements of other scientists and in general the scientific community he or she is a part of.

  23. Similarly to fn. 21, one may worry that the present account overgeneralizes what may count as collective knowledge such that any case of interpersonal interaction may count as a case of collective knowledge. In response, each time, we need to ask the following questions: First, judging by the criterion of continuous reciprocal interactivity, is the relevant process a collective belief forming process? And, second, is the output of the relevant process a true belief? So, for example, asking for directions from a stranger and acquiring a true belief on that basis won’t count as collective knowledge, but merely as individual, testimonial knowledge, because the relevant process is merely a one-way, linear interpersonal interaction. In contrast, as I have previously noted elsewhere (Carter et al. 2014) the products of transactive memory processes (Wegner et al. 1985; Wegner 1986)—provided that they are true—are typical cases of collective knowledge.

  24. Unless the overall belief-forming process of the research team can somehow be technologically emulated.

  25. Thanks to two anonymous referees for Synthese, whose insightful comments helped clarify the points laid out in the above two paragraphs.

  26. Consider Goldman (2004) attempt to explain Sandy Berger’s (former US national security adviser) mysterious dictum that “the F.B.I. didn’t know what it did know” about the 9/11 attack. Goldman claims that Berger refers to two different conceptions of the bureau. Under a summative conception of the agency as an aggregate of individual members, FBI did have knowledge of the attack, because certain agents individually possessed sufficient evidence, such that were this evidence to be put together, it would lead to a successful prediction of the attack. These individual agents, however, failed to appropriately combine their pieces of evidence in the right way meaning that the FBI, taken as a corporate or group entity, failed to know that the attack would take place.

  27. It is important to note, however, that, usually, philosophical zombies are not invoked to demonstrate that there can be minds that lack phenomenal consciousness. Their main role in the literature has been to demonstrate the shortcomings of physicalism: that is, that there can be creatures that are physically and behaviorally identical to us, but which nevertheless lack consciousness. Since, however, the standard description is that zombies “behave just like us, and some even spend a lot of time discussing consciousness” (Kirk 2011), the assumption seems to be that zombies may well qualify as mindful, even though they are entirely void of consciousness.

  28. One possible worry with the idea of minimal mindedness is that it may run the risk of being too liberal, depending on what may count as a psychological process or ability. At this point, employing a rather common tactic, Wilson responds without offering a definition. Instead, he attempts to ‘fix our ideas’ on the basis of the following incomplete but suggestive list of what may count as a psychological process or ability: “perception, memory, imagination (classical Faculties); attention, motivation, consciousness, decision-making, problem-solving (processes or abilities that are the focus of much contemporary work in the cognitive sciences); and believing, desiring, intending, trying, willing, fearing, and hoping (common, folk psychological states)” (Wilson 2001b, p. 266). Of course, it might still be objected that this is not an entirely successful approach to clarifying what may count as minimally mindful, as some of the most basic psychological processes and abilities listed above can be plausibly ascribed to systems whose mentality is rather dubious (for example, problem-solving or perceiving can be plausibly, even if metaphorically, ascribed to certain computers). Note, however, that this worry gradually fades away as we focus on more complicated psychological processes and abilities, such as imagining, decision-making, believing and, as in the present case, being justified. Surely, any system that posses such complicated psychological traits should qualify as at least minimally mindful.

  29. Even though this claim may sound as presupposing substance metaphysics, the present paper draws on dynamical systems theory (see Sect. 2.1), which should be a good indication that the present approach is rather sympathetic to the spirit and methodology of process philosophy (for an overview of the debate between substance metaphysics and process philosophy, see Seibt 2012). According to dynamical systems theory, however, properties cannot be conceptualized in the absence of the systems they belong to: Properties are behavioral regularities that arise out of processes of interactions between the components of a system. Accordingly, the present approach seems to be orthogonal to the debate between substance metaphysics and process philosophy as it does not ascribe metaphysical primacy to either substances or processes—an adequate description of reality requires both.

  30. Perhaps one way to do so is to follow Wilson (2001a, b, 2004, 2005) who has attempted to propose such a deflated approach to social properties on the basis of what he calls the ‘social manifestation’ thesis: Collective psychological properties, whatever they are, are properties of individuals, no matter they can only be manifested insofar the relevant individuals constitute part of a social group. The problem, however, is that Wilson’s insistence on the bearers of such collective properties being exclusively individuals is a form of favoritism towards individualism. Specifically, Wilson derives his social manifestation thesis from what he calls (a) wide and (b) radically wide realization: Respectively, the ideas that the total (in the case of (a)) and core (in the case of (b)) realization of certain properties is at least partly located outside the individual who possesses the relevant property (2001a). However, when Wilson is pressed to explain why the bearers of such properties are not themselves wide but are, instead, exclusively individuals, all he offers by way of an explanation are a few remarks to the effect that either the core realizers of the relevant property are realized in large part—even if not wholly—by the activity of the individual, or that individuals, in general, must stand out in our explanations, because they “are spatio-temporally bounded, relatively cohesive, unified entities that are continuous across space and time” (Wilson 2001a, p. 24). Since, however, in the case of distributed cognition the relevant collective properties are radically wide, realized by the activity of all the contributing individuals operating in tandem, and since distributed cognitive systems are spatio-temporally bounded, cohesive, unified wholes in themselves, none of these explanations is satisfactory. In fact, even Wilson himself seems to admit as much: “In at least some cases of wide realization, particularly those of radically wide realization, there is [no] non-arbitrary way to single out individuals as the subjects or ‘owners’ of the corresponding mental properties. If we have wide realizations of mental states, and thus wide mental states, so too we should have ‘wide subjects’ of those states” (ibid., p. 24).

  31. Alternatively, if one does not agree with this claim, it is possible to explain why lab technicians fail to form proper parts of the epistemic subject that deserves epistemic credit for the final piece of knowledge—despite qualifying as proper parts of the cognitive systems that gave rise to it—by complementing the above virtue reliabilist approach with the collective intentionality approach to group knowledge (Gilbert 2007a, b, 2010; List and Pettit 2006; List 2011; Rolin 2008): The lab technicians are not jointly committed to accepting or endorsing the result of the experiment and so they are not epistemically responsible or creditable for it. It should be noted, however, that, at least on certain formulations (e.g., Rolin 2008), the collective intentions approach to collective knowledge gives rise to collective epistemic responsibility in the robust sense of having the capacity to monitor and adjust one’s epistemic states in accordance with norms of rationality. In other words, at least certain formulations of the collective intentionality approach to group knowledge may be epistemically internalist in spirit and as such they may run counter the spirit of virtue reliabilism, which is an epistemically externalist approach to both individual and group knowledge. Thanks to an anonymous referee for suggesting this alternative as well as for proposing the thought experiment and objection discussed in the above three paragraphs.

  32. The collective intentionality approach to collective knowledge (Gilbert 2007a, b, c, 2010; Rolin 2008; List 2011; Wray 2007; Tuomela 2004) is not here considered as an alternative, since it does not focus on knowledge that is collectively produced, but only on knowledge being collectively possessed. But, even if there were a story for such an approach to tell with respect to knowledge that is collectively produced, there would still be way more to tell about how such a story would fit within contemporary epistemology.

References

  • Adams, F., & Aizawa, K. (2001). The bounds of cognition. Philosophical Psychology, 14(1), 43–64. doi:10.1080/09515080120033571.

    Article  Google Scholar 

  • Adams, F., & Aizawa, K. (2010). The bounds of cognition (1st ed.). Malden, MA: Wiley-Blackwell.

    Book  Google Scholar 

  • Bach-y-Rita, P., & Kercel, S. W. (2003). Sensory substitution and the human-machine interface. Trends in Cognitive Science, 7(12), 541–6.

    Article  Google Scholar 

  • Baehr, J. (2006). Character, reliability and virtue epistemology. The Philosophical Quarterly, 56(223), 193–212. doi:10.1111/j.1467-9213.2006.00437.x.

    Article  Google Scholar 

  • Bamberger, J. S., & Brofsky, H. (1979). The art of listening: Developing musical perception. New York: Harper & Row.

    Google Scholar 

  • Barnes, B., Bloor, D., & Henry, J. (1996). Scientific knowledge: A sociological analysis. Edinburgh: A&C Black.

    Google Scholar 

  • Barnier, A. J., Sutton, J., Harris, C. B., & Wilson, R. A. (2008). A conceptual and empirical framework for the social distribution of cognition: The case of memory. Cognitive Systems Research, 9(1–2), 33–51.

    Article  Google Scholar 

  • Berk, L., & Garvin, R. (1984). Development of private speech among low-income appalachian children. Developmental Psychology, 20(2), 271–286.

    Article  Google Scholar 

  • Block, N. (2002). Concepts of consciousness. In D. Chalmers (Ed.), Philosophy of mind: Classical and contemporary readings (pp. 206–219). Oxford: Oxford University Press.

    Google Scholar 

  • Bloomfield, L. (1933). Language. New York: Holt, Rinehart and Winston.

    Google Scholar 

  • Bloor, D. (1991). Knowledge and social imagery. Chicago: University of Chicago Press.

    Google Scholar 

  • Braddon-Mitchell, D., & Jackson, F. (2006). Philosophy of mind and cognition: An introduction (2nd ed.). Malden, MA: Wiley-Blackwell.

    Google Scholar 

  • Brewer, F.W. & Lambert, B.L. (2001). The theory ladeness of observation and the theory-ladeness of the rest of the scientific process. Philosophy of Science, Vol. 68, No. 3, Supplement: Proceedings of the 2000 Biennial Meeting of the Philosophy of Science Association. Part I: Contributed Papers, (pp. S176–S186).

  • Burge, T. (1986). Individualism and psychology. Philosophical Review, 95, 3–45.

    Article  Google Scholar 

  • Carter, J. A., Kallestrup, J., Palermos, S. O., & Pritchard, D. (2014). Varieties of externalism. Philosophical Issues, 24(1), 63–109.

    Article  Google Scholar 

  • Chalmers, D. (1996). The conscious mind. Oxford: Oxford University Press.

    Google Scholar 

  • Chemero, A. (2009). Radical embodied cognitive science. Cambridge: MIT press.

    Google Scholar 

  • Chemla, P. K. (2012). The history of mathematical proof in ancient traditions. Cambridge: Cambridge University Press.

    Google Scholar 

  • Churchland, P. M. (1979). Scientific realism and the plasticity of the mind. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Churchland, P. M. (1988). Perceptual plasticity and theoretical neutrality: A reply to Jerry Fodor. Philosophy of Science, 55, 167–187.

    Article  Google Scholar 

  • Churchland, P. M. (1989). A neurocomputational perspective. The nature of mind and the structure of science. Cambridge: The MIT Press.

    Google Scholar 

  • Clark, A. (1998). Magic words, how language augments human computation. In P. Carruthers & J. Boucher (Eds.), Language and thought: Interdisciplinary themes. Cambridge: Cambridge University Press.

    Google Scholar 

  • Clark, A., & Chalmers, D. (1998). The extended mind. Ananlysis, 58(1), 7–19.

    Article  Google Scholar 

  • Clark, A. (2007). Curing cognitive hiccups: A defense of the extended mind. The Journal of Philosophy, 104, 163–192.

    Article  Google Scholar 

  • Clark, A. (2008). Supersizing the mind. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Estany, A. (2001). The thesis of theory-laden observation in the light of cognitive psychology. Philosophy of Science, 68(2), 203–217.

    Article  Google Scholar 

  • Feyerabend, P. K. (1975). Against method: Outline of an anarchistic theory of knowledge. New York: New Left Books.

    Google Scholar 

  • Fodor, J. (1984). Observation reconsidered. Philosophy of Science, 51, 23–43.

    Article  Google Scholar 

  • Fodor, J. (1988). A reply to Churchland’s ‘perceptual plasticity and theoretical neutrality. Philosophy of Science, 55, 188–198.

    Article  Google Scholar 

  • Froese, T., Gershenson, C., & Rosenblueth, D.A. (2013). The dynamically extended mind. Retrieved http://arxiv.org/abs/1305.1958.

  • Fuller, S. (2007). The knowledge book: Key concepts in philosophy, science and culture. Durham: Acumen.

    Google Scholar 

  • Fuller, S. (2012). Social epistemology: A quarter-century itinerary. Social Epistemology, 26(3–4), 267–83.

    Article  Google Scholar 

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

    Book  Google Scholar 

  • Giere, R. (2002a). Discussion note: Distributed cognition in epistemic cultures. Philosophy of Science, 69, 637–644.

    Article  Google Scholar 

  • Giere, R. (2002b). Scientific cognition as distributed cognition. In Peter Carruthers, Stephen Stitch, & Michael Siegal (Eds.), Cognitive bases of science. Cambridge: Cambridge University Press.

    Google Scholar 

  • Giere, R., & Moffat, B. (2003). Distributed cognition: Where the cognitive and the social merge. Social Studies of Science, 33/2, 1–10.

    Google Scholar 

  • Giere, R. (2006). The role of agency in distributed cognitive systems. Philosophy of Science, 73, 710–719.

    Article  Google Scholar 

  • Giere, R. (2007). Distributed cognition without distributed knowing. Social Epistemology, 21(3), 313–320.

    Article  Google Scholar 

  • Gilbert, M. (2007c). Remarks on collective belief. Socializing Epistemology: The Social Dimensions of Knowledge 1994. Retrieved at SSRN: http://ssrn.com/abstract=1052361.

  • Gilbert, M. (2007a). Collective epistemology. Episteme, 1(2), 95–107. doi:10.3366/epi.2004.1.2.95.

    Article  Google Scholar 

  • Gilbert, M. (2007b). Modeling collective belief. Synthese, 73, 185–204.

    Article  Google Scholar 

  • Gilbert, M. (2010). Belief and acceptance as features of groups. Protosociology: An International Journal of Interdisciplinary Research, 16, 35–69.

    Google Scholar 

  • Goldman, A. (2004). Group knowledge versus group rationality: Two approaches to social epistemology. Episteme, 1(01), 11–22.

    Article  Google Scholar 

  • Goldman, A. (2010). Why social epistemology is real epistemology. In Adrian Haddock, Alan Millar, & Duncan Pritchard (Eds.), Social epistemology (pp. 1–28). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Goody, J. (1977). The domestication of the savage mind. Cambridge: Cambridge University Press.

    Google Scholar 

  • Goody, J. (1986). The logic of writing and the organization of society. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Goody, J. (1987). The interface between the oral and the written. Cambridge: Cambridge University Press.

    Google Scholar 

  • Greco, J. (1999). Agent reliabilism. In James Tomberlin (Ed.), Philosophical perspectives 13: Epistemology (pp. 273–296). Atascadero, CA: Ridgeview Press.

    Google Scholar 

  • Greco, J., & Zagzebski, L. (2000). Two kinds of intellectual virtue. Philosophy and Phenomenological Research, 60(1), 179. doi:10.2307/2653438.

    Article  Google Scholar 

  • Greco, J. (2004). Knowledge as credit for true belief. In M. DePaul & L. Zagzebski (Eds.), Intellectual virtue: Perspectives from ethics and epistemology. Oxford: Oxford University Press.

    Google Scholar 

  • Greco, J. (2007). The nature of ability and the purpose of knowledge. Philosophical Issues, 17, 57–69.

    Article  Google Scholar 

  • Greco, J. (2010). Achieving knowledge: A virtue-theoretic account of epistemic normativity. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Hanson, N. R. (1961). Patterns of discovery. Cambridge: Cambridge University Press.

    Google Scholar 

  • Hanson, N. R. (1969). Perception and discovery; An introduction to scientific inquiry. San Francisco: Freeman, Cooper.

    Google Scholar 

  • Harris, R. (1989). How does writing restructure thought? Language & Communication, 9(2–3), 99–106.

    Article  Google Scholar 

  • Haugeland, J. (1993). Mind embodied and embedded. In Y.-H. H. Houng, J. Ho (Eds.), Mind and cognition: 1993 international symposium (pp. 233–267) Academica Sinica.

  • Haugeland, J. (2000). Having thought: Essays in the metaphysics of mind (New ed.). Cambridge: Harvard University Press.

    Google Scholar 

  • Hempel, C. (1966). Philosophy of natural science. Englewood Cliffs, NJ: Prentice Hall.

    Google Scholar 

  • Hempel, C. (1970). Aspects of scientific explanation. New York: The Free Press.

    Google Scholar 

  • Heylighen, F., Heath, M., & Van Overwalle, F. (2007). The emergence of distributed cognition: A conceptual framework. In Proceedings of collective intentionality IV (2004), Vol. IV. University of Siena.

  • Hutchins, E. (1995). Cognition in the wild. Cambridge: MIT Press.

    Google Scholar 

  • Kaiser, D. (2005). Physics and Feynman’s diagrams. American Scientist, 93(2), 156.

    Article  Google Scholar 

  • Kirk, R. (2011). Zombies. The Stanford Encyclopedia of Philosophy.

  • Klein, U. (1999). Techniques of modeling and paper-tools in classical chemistry. In M. Morgan & M. Morrison (Eds.), Models as mediators: Perspectives on the natural and social science. Cambridge: Cambridge University Press.

    Google Scholar 

  • Knorr-Cetina, K. (1999). Epistemic cultures: How the sciences make knowledge. Cambridge: Harvard University Press.

    Google Scholar 

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

    Google Scholar 

  • Lakatos, I. (1970). Falsification and the methodology of scientific research programmes. In Imre Lakatos Alan Musgrave (Ed.), Criticism and the growth of knowledge. Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Latour, B., & Woolgar, S. (1986). Laboratory life: The construction of scientific facts. Princeton: Princeton University Press.

    Google Scholar 

  • Latour, B. (1999). Pandora’s hope: An essay on the reality of science studies. Cambridge, Mass: Harvard University Press.

    Google Scholar 

  • Latour, B. (2007). Reassembling the social: An introduction to actor-network-theory (New Ed ed.). Oxford: OUP Oxford.

    Google Scholar 

  • Levine, J. (1983). Materialism and qualia: The explanatory gap. Pacific Philosophical Quarterly, 64, 354–361.

    Google Scholar 

  • Linell, P. (2005). The written language bias in linguistics. London: Routledge.

    Book  Google Scholar 

  • List, C., & Pettit, P. (2006). Group agency and supervenience. The Southern Journal of Philosophy, 44(S1), 85–105. doi:10.1111/j.2041-6962.2006.tb00032.x.

    Article  Google Scholar 

  • List, C. (2011). Group agency: The possibility, design, and status of corporate agents. Oxford: OUP Oxford.

    Book  Google Scholar 

  • Logan, K.R. (2003). The extended mind: Understanding language and thought in terms of complexity and Chaos Theory. In D. McArthur & C. Mulvihil (Eds.) Humanity and the Cosmos. Retrieved http://www.upscale.utoronto.ca/PVB/Logan/Extended/Extended.html.

  • Logan, K. R. (2006). The extended mind model of the origin of language and culture. In N. Gontier, et al. (Eds.), Evolutionary epistemology and culture (pp. 149–167). New York: Springer.

    Chapter  Google Scholar 

  • Logan, K. R. (2008). The extended mind: The emergence of language, the human mind and culture. Toronto: University of Toronto Press.

    Google Scholar 

  • Markie, P. (2013). Rationalism vs. empiricism. In: E.N. Zalta (Ed.), The stanford encyclopedia of philosophy (Summer 2013.) Retrieved http://plato.stanford.edu/archives/sum2013/entries/rationalism-empiricism/.

  • McClelland et al. (1986). In J.L. McClelland, D.E. Rumelhart & the PDP Research Group (Eds.), Parallel distributed processing : Explorations in the microstructure of cognition, vol. 2. Cambridge, MA: MIT Press.

  • McGinn, C. (1989). Can we solve the mind-body problem? Mind, 98, 349–66.

    Article  Google Scholar 

  • McGinn, C. (1991). The problem of consciousness. Oxford: Blackwell.

    Google Scholar 

  • Menary, R. (2006). Attacking the bounds of cognition. Philosophical Psychology, 19(3), 329–344.

    Article  Google Scholar 

  • Menary, R. (2007). Cognitive integration: Mind and cognition unbound. Basingstoke: Palgrave McMillan.

    Book  Google Scholar 

  • Menary, R. (2013). Cognitive integration, enculturated cognition and the socially extended mind. Cognitive Systems Research, 25–26, 26–34. doi:10.1016/j.cogsys.2013.05.002.

    Article  Google Scholar 

  • Morgan, M., & Morrison, M. (1999). Models as mediators: Perspectives on the natural and social science. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Olson, D. R. (1996a). Language and literacy: What writing does to language and mind. Annual Review of Applied Linguistics, 16, 3–13.

    Article  Google Scholar 

  • Olson, D. R. (1996b). The world on paper: The conceptual and cognitive implications of writing and reading. Cambridge: Cambridge University Press.

    Google Scholar 

  • Olson, D. R. (2002). What writing does to the mind. In E. Amsel & J. P. Byrnes (Eds.), Language, literacy, and cognitive development: The development and consequences of symbolic communication. New York: Psychology Press.

    Google Scholar 

  • Palermos, S. O. (2011). Belief-forming processes, extended. Review of Philosophy and Psychology, 2, 741–765.

    Article  Google Scholar 

  • Palermos, S.O. (2013). Could reliability naturally imply safety?. European Journal of Philosophy. doi:10.1111/ejop.12046.

  • Palermos, S. O., & Pritchard, D. (2013). Extended knowledge and social epistemology. Social Epistemology Review and Reply Collective, 2(8), 105–120.

    Google Scholar 

  • Palermos, S. O. (2014a). Loops, constitution, and cognitive extension. Cognitive Systems Research, 27, 25–41.

    Article  Google Scholar 

  • Palermos, S. O. (2014). Knowledge and cognitive integration. Synthese, 191(8), 1931–1951.

    Article  Google Scholar 

  • Pettit, P. (2002). Collective reasons and powers. Legal Theory, 8(04), 443–470.

    Article  Google Scholar 

  • Popper, K. R. (1968). Epistemology without a knowing subject. In B. van Rootselaar & J. F. Staal (Eds.), Proceedings of the third international congress for logic, methodology and philosophy of science. Amsterdam. (Reprinted in K. R. Popper. (1972). Objective knowledge: An evolutionary approach. Oxford: Oxford University Press).

  • Pritchard, D. (2010b). Cognitive ability and the extended cognition thesis. Synthese, 175, 133–151.

    Article  Google Scholar 

  • Pritchard, D. (2010a). Knowledge and understanding. In A. Haddock, A. Millar, & D. H. Pritchard (Eds.), The nature and value of knowledge. Three Investigations, Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Putnam, H. (1975). The meaning of “Meaning”. In K. Gunderson (Ed.), Language, mind and knowledge. Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Rolin, K. (2008). Science as collective knowledge. Cognitive Systems Research, 9(1–2), 115–124. doi:10.1016/j.cogsys.2007.07.007.

    Article  Google Scholar 

  • Rowlands, M. (1999). The body in mind: Understanding cognitive processes. New York: Cambridge University Press.

    Book  Google Scholar 

  • Rupert, R. D. (2004). Challenges to the hypothesis of extended cognition. Journal of Philosophy, 101(8), 389–428.

    Article  Google Scholar 

  • Rupert, R. D. (2009). Cognitive systems and the extended mind (1st ed.). Oxford: OUP USA.

    Book  Google Scholar 

  • Saussure, F. (1916/1983). Course in general Linguistics. Duckworth. (Roy Harris, Trans.).

  • Seibt, J. (2012). Process Philosophy. In E. N. Zalta (Ed.), The stanford encyclopedia of philosophy (Fall 2013 Edition). Retrieved http://plato.stanford.edu/archives/fall2013/entries/process-philosophy/.

  • Shani, I. (2013). Making it mental: In search for the golden mean of the extended cognition controversy. Phenomenology and the Cognitive Sciences, 12(1), 1–26.

    Article  Google Scholar 

  • Shankweiler, D., & Liberman, I. (1972). Misreading: A search for causes. In J. Kavanagh & I. Mattingly (Eds.), Language by ear and by eye; the relationships between speech and reading (pp. 293–317). Cambridge, MA: MIT Press.

    Google Scholar 

  • Sosa, E. (1988). Beyond skepticism, to the best of our knowledge. Mind, New Series, 97(386), 153–188.

    Article  Google Scholar 

  • Sosa, E. (1993). Proper functionalism and virtue epistemology. Nous, 27(1), 51–65.

    Article  Google Scholar 

  • Sosa, E. (2007). A virtue epistemology: Apt belief and reflective knowledge. Oxford: Clarendon Press.

    Book  Google Scholar 

  • Suppe, F. (1977). The structure of scientific theories. Chicago: University of Illinois Press.

    Google Scholar 

  • Suppes, O. (1961). A comparison of the meaning and use of models in the mathematical and empirical sciences. In H. Freudenthal (Ed.), The concept and role of the model in mathematics and natural and social sciences (pp. 163–177). Dordrecht: Reidel.

    Chapter  Google Scholar 

  • Suppes, O. (1967). What is a scientific theory? In S. Morgenbesser (Ed.), Philosophy of science today (pp. 55–67). New York: Basic Books.

    Google Scholar 

  • Sutton, J., Barnier, A., Harris, C., & Wilson, R. (2008). A conceptual and empirical framework for the social distribution of cognition: The case of memory. Cognitive Systems Research, 9(1—-2), 35–51.

    Google Scholar 

  • Sutton, J. (2008). Between individual and collective memory: Coordination, interaction, distribution. Social Research, 75(1), 23–48.

    Google Scholar 

  • Theiner, G. (2013). Language, literacy, and media theory: Exploring the cultural history of the extended mind. AVANT. Journal of the Philosophical-Interdisciplinary Vanguard, 4(2).

  • Theiner, G., & Wilson, R. (2013). Group mind. Encyclopedia of Philosophy and the Social Sciences.

  • Theiner, G., Allen, C., & Goldstone, R. (2010). Recognizing group cognition. Cognitive Systems Research, 11(4), 378–395.

    Article  Google Scholar 

  • Theiner, G., & O’Connor, T. (2010). The emergence of group cognition. In A. Corradini & T. O’Connor (Eds.), Emergence in science and philosophy (pp. 6–78). London: Routledge.

    Google Scholar 

  • Theiner, G. (2011). Res cogitans extensa: A philosophical defense of the extended mind thesis. Bern: Peter Lang GmbH, Europaischer Verlag der Wissenschaften.

    Google Scholar 

  • Tollefsen, D. (2004). Collective intentionality. Internet Encyclopedia of Philosophy.

  • Tollefsen, D. (2002a). Organizations as true believers. Journal of Social Philosophy, 33(3), 395–410.

    Article  Google Scholar 

  • Tollefsen, D. (2002b). Collective intentionality and the social sciences. Philosophy of the Social Sciences, 32(1), 25–50. doi:10.1177/004839310203200102.

    Article  Google Scholar 

  • Tollefsen, D. (2006). From extended mind to collective mind. Cognitive Systems Research, 7(2–3), 140–150.

    Article  Google Scholar 

  • Tollefsen, D., & Dale, R. (2011). Naturalizing joint action: A process-based approach. Philosophical Psychology, 25, 385–407.

    Article  Google Scholar 

  • Tomasello, M. (2009). The cultural origins of human cognition. Cambridge: Harvard University Press.

    Google Scholar 

  • Tufte, E. R. (2001). The visual display of quantitative information (2nd ed.). Cheshire, Conn: Graphics Press USA.

    Google Scholar 

  • Tuomela, R., & Miller, K. (1988). We-intentions. Philosophical Studies, 53(3), 367–389. doi:10.1007/BF00353512.

    Article  Google Scholar 

  • Tuomela, R. (2004). Group knowledge analyzed. Episteme, 1(2), 109–127.

    Article  Google Scholar 

  • van Fraassen, B. (1980). The scientific image. Oxford: Clarendon Press.

    Book  Google Scholar 

  • van Gelder, T. (1995). What might cognition be if not computation? Journal of Philosophy, 92(7), 81–345.

    Article  Google Scholar 

  • Van Gulick, R. (2004). Consciousness. Stanford Encyclopedia of Philosophy.

  • Vygotsky, L. (1978). Mind in society: Development of higher psychological processes (New ed.). Cambridge: Harvard University Press.

    Google Scholar 

  • Vygotsky, L. (1986). Thought and language (2 Revised ed.). Cambridge, Mass: MIT Press.

    Google Scholar 

  • Wegner, D., Giuliano, T., & Hertel, P. (1985). Cognitive interdependence in close relationships. In W. J. Ickes (Ed.), Compatible and incompatible relationships (pp. 253–276). New York: Springer-Verlag.

    Chapter  Google Scholar 

  • Wegner, D. (1986). Transactive memory: A contemporary analysis of the group mind. In B. Mullen & G. R. Goethals (Eds.), Theories of group behavior. New York: Springer-Verlag.

    Google Scholar 

  • Weiskopf, D. A. (2007). Concept empiricism and the vehicles of thought. Journal of Consciousness Studies, 14(s 9–10), 156–183.

    Google Scholar 

  • Wertsch, J. V. (1988). Vygotsky and the social formation of mind (Reprint ed.). Cambridge, Mass.: Harvard University Press.

    Google Scholar 

  • Wheeler, M. (2005). Reconstructing the cognitive world. Cambridge, MA: MIT Press.

    Google Scholar 

  • Wilson, R. A. (2000). The mind beyond itself. In D. Sperber (Ed.), Metarepresentations: A multidisciplinary perspective (pp. 31–52). New York: Oxford University Press.

  • Wilson, R. (2001a). Two views of realization. Philosophical Studies, 104, 1–31.

    Article  Google Scholar 

  • Wilson, R. (2001b). Group-level cognition. Philosophy of Science, 68(3), 262–273.

    Article  Google Scholar 

  • Wilson, R. (2004). Boundaries of the mind: The individual in the fragile sciences: Cognition. New York: Cambridge University Press.

    Book  Google Scholar 

  • Wilson, R. (2005). Collective memory, group minds, and the extended mind thesis. Cognitive Processing, 6(4), 227–236.

    Article  Google Scholar 

  • Wray, K. B. (2007). Who has scientific knowledge? Social Epistemology, 21(3), 337–347. doi:10.1080/02691720701674288.

    Article  Google Scholar 

  • Zagzebski, L. T. (1996). Virtues of the mind: An inquiry into the nature of virtue and the ethical foundations of knowledge. New York, NY: Cambridge University Press.

    Book  Google Scholar 

Download references

Acknowledgments

I am grateful to J. Adam Carter and John Greco for very useful discussions on the topic and feedback on previous drafts. I am also thankful to two anonymous referees for Synthese for their particularly insightful reviews. Research into the area of this paper was supported by the AHRC-funded ‘Extended Knowledge’ Project, based at the Eidyn research centre, University of Edinburgh.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Spyridon Orestis Palermos.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Palermos, S.O. Active externalism, virtue reliabilism and scientific knowledge. Synthese 192, 2955–2986 (2015). https://doi.org/10.1007/s11229-015-0695-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-015-0695-3

Keywords

Navigation