zammito - kant´s teleology then and now

23
5/21/2018 Zammito-Kant´sTeleologyThenandNow-slidepdf.com http://slidepdf.com/reader/full/zammito-kants-teleology-then-and-now 1/23 Teleology then and now: The question of Kant’s relevance for contemporary controversies over function in biology John Zammito Department of History, MS42, PO Box 1892, Rice University, Houston, TX 77005, USA Abstract ‘Naturalism’ is the aspiration of contemporary philosophy of biology, and Kant simply  cannot be refashioned into a naturalist. Instead, epistemological ‘deflation’ was the decisive feature of Kant’s treatment of the ‘biomedical’ science in his day, so it is not surprising that this might attract some philosophers of science to him today. A certain sense of  impasse in the contemporary ‘function talk’ seems to motivate renewed interest in Kant. Kant—drawing on his eighteenth-century predeces- sors—provided a discerning and powerful characterization of what biologists had to explain in organic form. His difference from the rest is that he opined that it was  impossible to explain it. Its ‘inscrutability’ was intrinsic. The third  Critique  essentially proposed the reduction of biology to a kind of pre-scientific descriptivism, doomed  never  to attain authentic scientificity, to have its ‘New- ton of the blade of grass’. By contrast, for Locke, and  a fortiori  for Buffon and his followers, ‘intrin- sic purposiveness’ was a fact of the matter about concrete biological phenomena; the features of internal self-regulation were hypotheses arising out of actual research practice. The difference comes most vividly to light once we recognize Kant’s distinction of the concept of  organism from the con- cept of  life. If biology must conceptualize self-organization as actual in the world, Kant’s regulative/ constitutive distinction is pointless in practice and the (naturalist) philosophy of biology has urgent work to undertake for which Kant turns out  not  to be very helpful.  2006 Elsevier Ltd. All rights reserved. Keywords:  Function; Immanuel Kant; Naturalism; Intrinsic purposiveness; Biology as a ‘special science’; Systematicity 1369-8486/$ - see front matter   2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.shpsc.2006.09.008 E-mail address:  [email protected] Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748–770 www.elsevier.com/locate/shpsc Studies in History and Philosophy of Biological and Biomedical Sciences

Upload: taonta

Post on 13-Oct-2015

43 views

Category:

Documents


1 download

DESCRIPTION

Kant and the problem of teleology

TRANSCRIPT

  • organic form. His dierence from the rest is that he opined that it was impossible to explain it. Itsinscrutability was intrinsic. The third Critique essentially proposed the reduction of biology to a

    Systematicity

    E-mail address: [email protected]

    Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770

    www.elsevier.com/locate/shpsc

    Studies in Historyand Philosophy ofBiological andBiomedical Sciences1369-8486/$ - see front matter 2006 Elsevier Ltd. All rights reserved.kind of pre-scientic descriptivism, doomed never to attain authentic scienticity, to have its New-ton of the blade of grass. By contrast, for Locke, and a fortiori for Buon and his followers, intrin-sic purposiveness was a fact of the matter about concrete biological phenomena; the features ofinternal self-regulation were hypotheses arising out of actual research practice. The dierence comesmost vividly to light once we recognize Kants distinction of the concept of organism from the con-cept of life. If biology must conceptualize self-organization as actual in the world, Kants regulative/constitutive distinction is pointless in practice and the (naturalist) philosophy of biology has urgentwork to undertake for which Kant turns out not to be very helpful. 2006 Elsevier Ltd. All rights reserved.

    Keywords: Function; Immanuel Kant; Naturalism; Intrinsic purposiveness; Biology as a special science;Teleology then and now: The questionof Kants relevance for contemporarycontroversies over function in biology

    John Zammito

    Department of History, MS42, PO Box 1892, Rice University, Houston, TX 77005, USA

    Abstract

    Naturalism is the aspiration of contemporary philosophy of biology, and Kant simply cannot berefashioned into a naturalist. Instead, epistemological deation was the decisive feature of Kantstreatment of the biomedical science in his day, so it is not surprising that this might attract somephilosophers of science to him today. A certain sense of impasse in the contemporary function talkseems to motivate renewed interest in Kant. Kantdrawing on his eighteenth-century predeces-sorsprovided a discerning and powerful characterization of what biologists had to explain indoi:10.1016/j.shpsc.2006.09.008

  • congejust this ambition for naturalism (Quine, 1969). The Churchlands and their train in philos-ophy of mind and cognitive science cherish a similar ambition (see, for example, Church-

    land, 1995). It should be acknowledged, as well, that someperhaps manypracticingm. The father of naturalized epistemology, Willard van Orman Quine, cherished1985; Wagner & Warner, 1993). A notable variant, indeed, could be taken to be the con-tinuation (by other means) of the positivist program of unity of science via physical reduc-

    tionisries of views that have starkly disparate premises and programs (Depew & Weber,The self-understanding of an empirical science, biology : : : is for biologists to decide.A philosopher can only analyze the metaphysical costs of the various options.(McLaughlin, 2001, p. 190)

    If a teleological point of view is a condition for biology, teleology should rather beseen as a constitutive principle for this science. (Quarfood, 2004, p. 119)

    This paper is an exercise simultaneously in presentism and historicism (see Zammito,2004b). It will begin with a consideration of present issues in the philosophy of biology,and with recent conjecture that it might be illuminating to go back to Kants discussion(Quarfood, 2004, p. 118). Three recent writings move from current conundrums in thephilosophical discussion of function, especially in biology, to the question of whether Kantmight be of renewed relevance. Most important for my purposes will be McLaughlin(2001). A close second is Lewens (2004 and Forthcoming). Third is Quarfood (2004).All three pieces express the viewwhich inspires this entire special issuethat Kants tel-eology is of rich relevance to the current debate on biological function.

    This presentist consideration will lead then to a historicist consideration of what Kantactually argued. My claim is that naturalism is the aspiration of contemporary philoso-phy of biology (Bedau, 1991, 1993, Depew & Weber, 1985), and Kant simply cannot berefashioned into a naturalist (see Zammito, 2003). I will argue, instead, that Kant is mostaccommodating to present thinkers who wish to set in epistemological suspension thequestion of the actuality of function in biology (e.g. Lewens, 2004, p. x). Thus, therecourse to Kant is not merely a historical retreat. Indeed, epistemological deation(from constitutive to regulative, that is, from explanatory to heuristic) was the decisivefeature of Kants treatment of the life sciences of his day. In that light it is not surprisingthat this might attract some philosophers of science to him today.

    Of course, no one has ever really forgotten Kants discourse on teleology, but it washardly central to the philosophical assault on design that characterized the epoch of pos-itivism in the philosophy of sciencean epoch that, if at all, only recently expired (Zam-mito, 2004a). In the rst part of that positivist era, the nineteenth century, the ambitionwas to purge biology of natural theology, or what today calls itself intelligent design.Darwins natural selection was the purgative. Later, in the twentieth century, the eortwas to purge philosophy of all metaphysics and to rene scientic explanation intoone denitive algorithmultimately, the so-called D-N model of logical empiricism.The upshot made teleology taboo in the physical sciences and an embarrassment in biol-ogy (and the social sciences). Because he held mechanistic methodology to be indispens-able for science, Kant could be comfortably assimilated to logical empiricism. In fact,Kant was held to be an esteemed progenitor of positivism in its sophisticated twentieth-century incarnationthough, like all venerated ancestors, best left at rest in his urn.

    One form of post-positivism, naturalism, has gradually captivated philosophers of sci-enceand especially of biology. Of course, naturalism is not just one form: it betokens a

    J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 749

    LeonardoHighlight

    LeonardoHighlight

    LeonardoHighlight

    LeonardoHighlight

    LeonardoHighlight

    LeonardoHighlight

    LeonardoHighlight

  • biologists believe they must still conform to a reductionism that is not only ontological butmethodological in order to be taken as authentic scientists. This suggests that the argu-ment for the disunity of science and the discrediting of the positivist myth of physicalreductionism need to be more eectively advocated in and for that community of scholars.For my part, I take this to be the appropriate naturalistic program, though I can hardlymake that case here (see Zammito, 2004a). In any event, I propose to assess Kants poten-tial assistance in terms of the pervasive commitment to naturalism in contemporary phi-losophy of biology.

    1. Conundrums of function talk in biology today

    A certain sense of impasse in the contemporary discourse seems to be the common moti-vation behind renewed interest in Kant (Rescher, 1986; Ruse, 2000; Walsh, 1996). Dead-ened by the dull thud of conicting intuitions (Bigelow & Pargetter, 1998, p. 257), manynd themselves asking whether there is no end to function talk in biology (Lewens, 2001).The same two basic alternatives have been debated back and forth over the past fortyyears (McLaughlin, 2001, p. 63). There is, in the paradoxical formulation of Godfrey-Smith (1993), a consensus without unity. This pluralism appears to be a truce of exhaus-tion, not a truth of conviction. But, as Lewens argues, important issues are hardlyresolved: how teleological approaches in biology work, what risks they carry, what formsof teleological content can be grounded by biological processes, and why teleologicalapproaches are found only in biological contexts (Lewens, 2004, p. 7). He draws the gistof conventional wisdom, that is, that natural selection : : : plays a role analogous to inten-tional choice and thus grounds various claims about function : : : in the natural world;moreover, selection can give [functional] traits norms that can be met, hence it canground projects to naturalize content in the philosophy of mind. But then he asserts:such a picture is almost completely mistaken (ibid, p. 3). His book goes on to make thatcase. Similarly, McLaughlin notes that a great deal of the interest in functional explana-tion is due to naturalistic projects, especially in the philosophy of mind (McLaughlin,2001, p. 10). Does natural selection as such get you all the teleology you need for a nat-uralistic interpretation of functional explanation? The answer will turn out to be no (ibid.,p. 14). And his book goes on to make that case.

    What are the conundrumsthe impassesof contemporary function talk, that Kantswork of the eighteenth century should be of relevance in coming to terms with them? Obvi-ously, one can only undertake a partial consideration of this vast literature. I wish to teaseout indicative, not exhaustive traits. At the most immediate level, the question of the anal-ogy of human artifacts with natural organisms presents itself, with a lineage that goes allthe way back to at least Aristotle. The phenomenal encounter with natural organisms elic-its the sense of parallel with objects of human design, and what so strongly impinges as adescriptive analogy invites translation into a causal explanation. Yet analogy implies notonly similarity but dierence. The imputation of design is metaphorical, and metaphorcannot be literal: organisms are not artifacts (Lewens, 2004, p. ix). Via disanalogy Aris-totle discerned a fundamental, ontological dierence between external and internal teleol-ogy. Kant, too, recognized both the analogy and the disanalogy, but for him theimplication was a retreat from explanatory perplexity to epistemological discretion, aswe will discuss below. For more recent commentators, the propensity to nd the analogy

    750 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770heuristically compelling and yet untting methodologically made an escape into ecient

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 751cause seem essential: metaphor needed to be attened out into the literal lines of standardcausation. And yet, unreduced teleology : : : surfaces time and again even in the most aus-terely naturalistic analyses (Quarfood, 2004, p. 157).

    The two approaches between which the debate has been ongoing for the last forty yearsare known as the etiological and the causal-role conceptions of function. The etiologicalapproach, as its name betokens, looks backward for an account of the origin or cause ofthe item to which function is ascribed, while the causal-role approach looks forward to theeect the item will have in serving a given systemic capacity. The two views can be tracedback to Carl Hempel and to Ernest Nagel, respectively, in the 1960s (McLaughlin, 2001,Ch. 46). As two of the foremost logicalempiricists, Hempel and Nagel sought toappraise their respective conceptions of function in terms of the general theory of expla-nation expressed by the D-N model. Hempel denied that function could satisfy the logicalrequirements for valid explanation, and Nagel rescued its validity by robbing it of its spec-icity (Nagel, 1977). The core of their endeavor was rendering teleological discourse with-out remainder into ecient cause explanation. That is, teleology was taken as a facon deparler, a metaphor or analogy, which could and should be cashed out by translation into asuitably logicalscientic explanation in terms of ecient cause: linear succession in timefrom cause to eect. This was assimilated entirely into the second round of interpretationcentered specically within the philosophy of biology, which entailed reduction of teleo-logical locutions to a naturalistically safe concept that : : : depends only on ecient cau-sation (Quarfood, 2004, p. 122). This more recent instauration was launched by Wright(1973), on the one side, and Cummins (1975), on the other.

    More recent thinkers in the tradition of Wrightsuch as the proper function theoristsand their modern history amenders (Millikan, 1989; Neander, 1991; Godfrey-Smith,1994)were inspired by the prospect that natural selection could provide the decisive e-cient-causal force to naturalize teleology. Function could be translated as the causalresult of a successful adaptation. Thus, for Sober (1993), function just is adaptation.The presence of function in a given item or trait token is the causal result of the contribu-tion of earlier tokens of the same type to the reproductive success of the organism. Becauseearlier tokens of the type establish its causal ecacy, the presence of the latest token is theoutcome of an ecient causal sequence, escaping the backward causality that made tel-eology unacceptable. The anthology, Natures purposes (Allen, Beko, & Lauder, 1998),presented an authoritative collection of key essays establishing this seeming consensus.The core consensus (Buller, 1999) was to explain in terms of ecient causes all the orderdiscerned in organic processes.

    The etiological view has seemed to loosen its grip in the last few years. Both Lewens andQuarfood indicate that revisionists have been gaining ground in casting suspicion on themodel. What seems to have come most under suspicion is the adaptationist interpretationof natural selection, especially its optimization model, and this primarily because of scru-ples about historical reconstruction. The start of the unraveling appears to have come withthe intervention of Gould and Lewontin (1979) against the Panglossian paradigm. It car-ried forward to a general recognition that it is very dicult on the basis of present adap-tiveness to infer both the environmental pressures and the adaptational sequences out ofwhich present organisms arose. As Lewens puts it, development plays as much of a role inthe explanation of adaptation as selection (Lewens, 2004, p. 16). If one cannot establishthe developmental constraints in the structure of the organism, one can never establish

    what selection pressures had to work with. As Lewens puts it, there is a question of

    LeonardoHighlight

  • 752 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770whether selection, or development, has the primary role in explaining form (ibid., p. 74).Moreover, there could be highly adaptive traits that were never adaptations, because theydid not vary across an entire population, and traits could become adaptive or cease to beadaptive as environments changed (Gould & Vrba, 1982). Indeed, it was clear that organ-isms could themselves alter their environments (Lewontin, 1985). All of this militatedskepticism about our abilities to uncover the evolutionary past (Lewens, 2004, p. 42).Lewens invokes Paul Griths for the conclusion: observed form underdetermines the nat-ure of the adaptive problems solved (ibid., p. 50; Griths, 1996). Thus, our hypothesesabout [an items] design history will often be underdetermined by those data, so that it isclearly a mistake to assume that we can always make a reliable inference to the best expla-nation when we attempt to infer past selection from the observations of organismic formalone (Lewens, 2004, pp. 5152). There may be many combinations of phenotype sets,heritability assumptions, tness measures and state equations compatible with the claimthat a phenotype is the ttest of available variants (ibid., p. 57). That is, not one but sev-eral histories can be constructed for the same trait (ibid., p. 58). Accordingly, there is morehistoricism to naturalism than the historical adaptationists understood. Of this there willbe more to say in my conclusion.

    At the same time the consideration began to arisefueled by a far wider disputebetween developmental and evolutionary biologiststhat the origin-orientation of theselectionists could not adequately account for the ongoing (physiological) self-regulationof systems in individual organisms (Griths and Gray, 1994). That is, evolution and phys-iology seem to entail quite dierent forms of function and require quite dierent accounts.The type/token distinction is of no use in the physiological context, and the question ofbackward causalityor better, of holistic causalitybecomes unavoidable. The notionthat the whole can be temporally prior to the parts and thus have a causal impact on thembrings up the problem of holistic causality (McLaughlin, 2001, pp. 2425). And jugglingtypes and tokens wont solve this problem, because we are caught up with the specictoken (ibid., p. 163).

    We need to understand each token organism as implicitly a feedback system in itself,reassembling [itself] all the time (ibid., p. 167). We have to nd some kind of feedbackmechanism that applies even to the rst generation of a new trait or organism : : : Itmay have to be explained by a number of detailed physiological processes (ibid., p.163). What characterizes these is a goal-directedness whose goal is ultimately the survivaland/or propagation of that system (ibid., p. 69). The basic assertion is that whatever hasan ergon can be the subject of benet. Anything that supports the characteristic activity ofan entity is good for that entity as an entity of that kind (ibid., p. 199). But it is not clearhow to make a compelling account of this. What seems required is a kind of historicallystretched quasiholistic causal relation : : : within one generation : : : [A] token of a type oftrait can be viewed as contributing to its own re-production as the same token : : : At eachparticular moment, the whole is determined by the properties of its parts : : : [W]e considerthe system to be the same : : : over time even if it consists of dierent parts at dierenttimes, and thus can have existed prior to some of its component parts (ibid., p. 210).Function interpreted in the sense of internal teleology : : : could not be (part of) the e-cient cause of the origin of the system because it can only exist once the system itself exists(ibid., p. 24). What is needed is a kind of system whose characteristic activity or ergon is toprovide for its own welfare (ibid., p. 76). That entails commitment to a complex-systems

    model of teleology (Christensen, 1996; Schlosser, 1998).

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 753Thus, immanent normativitywhat a system (and hence its functions) should or can besupposed to beis essential to the conceptualization (Bedau, 1992; Schurz, 2001; Wach-broit, 1994). But systems that satisfy this condition need not be living. In this context,the question becomes pressing whether self-regulating systems should really be restrictedto living organisms or might well need extension downward into physico-chemical pro-cesses of emergence, on the one hand, and upward into social systems, on the other. Thatthe actual order that self-regulation entails seems to require a notion of immanent norm-ativity seems starkly disconcerting for the positivist tradition of natural-scientic explana-tion, which enshrined the fact-value dichotomy as dogma (Putnam, 2002). It is awkwardeven for some versions of naturalism, which aim precisely to reduce the seemingly evalu-ativeintentional aspect of humans to materialecient causality, crudely: mind to brain(Wagner, 1996; Millikan, 1997). Within the teleosemantic project in philosophy of mind,McLaughlin notes, the concept of norm : : : is, of course, technical not moral: : : talkingabout norms of production, not of moral evaluation (McLaughlin, 2001, p. 104). Butfunction talk cannot get away from the question of welfare or benet, and the questionof naturalizing such notions is central to the current conundrums.

    All this brings back with a vengeance the original tension between artifacts and organ-isms, between external and intrinsic purposiveness, extending the question decisively to theactual nature of humannot simply organismicspontaneity and systematicity. For nat-uralism, McLaughlin puts it precisely, the explanatory relation between natural and arti-factual functions must be one of the generation of the latter from the former not of mutualsubsumption under a generic term (McLaughlin, 2001, p. 15). That is, naturalism is anontological historicism, that is, its project is to reconstruct the generation of artifactualfunctions from natural ones (ibid.). Humans are organisms rst, and designers only deriv-atively. Design is ontologically parasitical upon a designer: there can be no relative purpo-siveness without an intrinsic purpose to be served. The analogy of design presumes thetransparencyboth epistemological and ontologicalof human agency.

    Aristotelian internal teleology intrudes as an ontological matter. Lewens lays out theargument. It is the internal constitution of biological items, not the fact that selection actsonly on biological items, that best explains the appearance of artifact talk in biology alone(Lewens, 2004, p. 3). Yet one makes a mistake if : : : one assumes that there must be somespecic process, which only organisms undergo, that bestows normative, purposive stateson them (ibid., p. 120). I take this to mean there are non-organismic systems that qualifyas well. The basic point to be made is that only certain kinds of systems (systems whoseexact characterization I leave to others) are able to evolve by natural selection to yieldcomplex adaptations, Lewens writes (ibid., p. 125). Yet just this unwillingness to go fur-ther betokens his ultimately deationary project, aimed to expose the epistemologicalaporias of function talk: The problem in making a decisive choice between the theoriesis that there is really no single killer intuition that will tell us which of the accounts isright (ibid., p. 108). Lewens contents himself with demonstrating how a number of cru-cial disanalogies between selection and intention : : : can lead us to : : : underestimate thefunctional interconnectedness of organic, as opposed to articial, design (ibid., pp. 1617). He is interested in diminishing what he calls heavy function talk, though he notesthat there are aspects of what he calls the agent modelthat is, intrinsic purposivenesswhich appear very pertinent. The study of development makes goal-directedness the mosttempting form for an account of function (Lewens, Forthcoming, p. 5). He nds that

    today such goal-based theories are highly fashionable (ibid., p. 19). Thus, the new turn

  • to seeking facts about the internal organization or development of organisms, to physi-ology and morphology, rather than historical origin (ibid., p. 3). But Lewens wants todownplay the seriousness of heavy function talk within biology (ibid., p. 18). He is notsure whether goal-directedness can be justied as anything more than anthropomorphicprojection (ibid., p. 19). He doubts whether the goal can be specied in a manner thatallows falsication, or even goal-failure. Thus Lewens turns to Kant, who recognizeswith him all the appeals of both the artifact and the agent models, but recognizes theirweaknesses as well, and insists that teleology can only be heuristic.

    Yet nding the inadequacies of the artifact model for organisms is not enough, as Lew-ens himself recognizes: the actuality of systems with intrinsic purposiveness has to be con-ceptualized. This internal purposiveness, as Peter McLaughlin put it, is where the analogybetween artifacts and organism breaks down (McLaughlin, 2001, p. 145). The artifactmodel and the analogy of design simply stop short of the mark. McLaughlin goes on, deci-sively: the reality of internal (nonintentional) teleology is what is really at issue (ibid., p.17). The real metaphysical cost of functional explanation lies in a commitment to the exis-tence of entities that can stop a functional regress (ibid., p. 211). That is, the real problemhas always been holism (ibid., p. 27). To deal with this ontological challenge, McLaughlininsists we need to reconsider a deep ambiguity in the reception of teleology from the Aris-totelian tradition, the blur of nal cause with formal cause. While teleology and functionhave always, especially on the basis of the artifact model or analogy to design, seemed amatter of nal cause, McLaughlin argues this is in fact inaccurate: most genuinely func-tional explanations involve not so much an illicit appeal to nal causes as an implicitappeal to holistic causality (ibid., p. 9).

    What, then, are the issues current function talk would carry back to a consideration ofKant? First, the question of the phenomenal encounter with organisms: how are they rec-ognized, and as what? Second, the question of the analogy of artifact and organism: howsatisfactory is this analogy for biology? Third, the question of the exhaustiveness of e-cient cause for scientic explanation: can a mechanical account of origin (natural selection)suce as an account of development and physiology (self-regulative systematicity)? Fourth,what is intrinsic purposiveness for biological science? Is it possible to conceptualize itadequately? Fifth, how can normativity be understood within a (naturalist) scienticresearch program?

    2. Kant and teleology

    The intentions informing Kants composition of the Critique of (the power of) judgmentare exceedingly complex and multifarious (Zammito, 1992). On the face of it, critics havealways found it arbitrary that Kant treated of aesthetics and of biology in the same work.That this was a critique of judgment throughout seemed an architectonic ruse, not a sub-stantive consideration, and that the seemingly discrete considerations of aesthetics and biol-ogy were in fact part of a still larger construction entailing the problem of a unied order ofnature as empirical law, on the one hand, and the problem of humanmoral teleology, on theother, made the third Critique only more of a morass for critics. Nonetheless, Kants over-arching concern was not merely architectonic. Reconciliation of the rst and second Cri-tiquesof theoretical with practical reasonentailed the critical reconciliation of humanobligation, and its implicated freedom, with the lawfulness of phenomenal nature, the unity

    754 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770of its order, as the locus in whichman was to act out this obligation. Everything in the third

    LeonardoHighlight

    LeonardoHighlight

  • Critique was a means to that end (Genova, 1975; Zammito, 1992). Paul Guyer is surely cor-rect tomaintain that the basic reason for discussing organisms at all was precisely that theseare objects within our experience that can prompt us to take this twofold view of naturerequired to achieve the reconciliation (Guyer, 2001, p. 262). It may be theoretical diculties

    J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 755in comprehending organisms that require us to conceive them as products of purpose, but: : : it is our morally grounded conception of our own purposiveness as free that leads us tothe further thought that purposiveness entails immateriality, thus that organisms and ulti-mately all of nature must have an immaterial ground (ibid., p. 280).

    That is, Kants treatment of biology was always subsidiary to larger systematic concernsof the critical philosophy as a whole. That was a sword with two edges, however. It wasmeant to enable the coherence of his system, but it could also turn out to pose it grave prob-lems. I have suggested that biology created acute diculties for the unity of science in acoherent order of nature along the lines Kant preferred (Zammito, 2003). He had tworecourses: rst, to reformulate the issues of biology as regulative rather than constitutive,that is, as descriptive (heuristic) rather than explanatory (scientic). That was the course heset in his philosophy of (biological) science. Second, as Guyer suggested, he could proposethe possibility of thinkingthough, of course, not knowinga supersensible ground thatmight rescue the coherence of that order (at a metaphysical cost, even as thought, beyondthe philosophical bankroll of any current naturalist). We have good historical reason tobelieve that Kant made a decided shift over time from participation in actual theorizingin life science (to be sure, from his armchair) to a much more skeptical critique of itsmethod. Phillip Sloan characterizes Kants trajectory as a shift from the new history ofnature [Naturgeschichte] associated with Buon and Bonnet back to a transcendentalphilosophical version of description of nature [Naturbeschreibung] (Sloan, 2006, thisissue; Lagier, 2004 and Adickes, 1924). The third Critique essentially proposed thereduction of life science to a kind of pre-scientic descriptivism, doomed never to attainauthentic scienticity, never to have its Newton of the blade of grass (Kant, 1910,Vol. 5, p. 400).1

    For his relevance to contemporary (naturalist) philosophy of biology, Marcel Quarfoodsets the discussion of Kant in the proper frame by asserting: The distinctive feature ofKants view is : : : an epistemic presupposition constitutive for the study of life, rather thana denite ontological commitment (Quarfood, 2004, p. 145). Joan Steigerwald agrees Kantwas concerned with the conditions of the possibility of our cognition of organisms, notwith their fundamental nature. She stresses that Kant claimed we could only grasp organ-isms through analogy with our own purposive activity, that is, he maintained only theanalogy to artice gave us conceptual access to organic form (Steigerwald, 2006, this issue).The question that Steigerwald brings to the center of consideration is how Kant conceivedintrinsic purposivenessthe capacity to self-organize and to maintain organization as nat-urally occurring processes in organic bodies.

    Kant opened his Critique of teleological judgment with a consideration of the struc-ture of birds regarding how their bones are hollow, how their wings are positioned to pro-duce motion and their tails to permit steering : : : (Kant, 1987, p. 236 [5:360]).2 The

    1 All translations are my own unless otherwise indicated.

    2 References in square brackets are to the original German in Kant (1910).

  • 756 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770features he highlighted are anatomicalphysiological, betting adaptiveness or causal-role thinking today. The question he immediately raised is how nature could have broughtthese about, that is, an etiological question. He asserted that there was no way that themere nexus eectivus in nature could account for this production; it was irretrievablycontingent. Famously, he claimed we resort to nexus nalis, causality according to pur-pose, in order to organize our reception of this phenomenon. But we do this only : : : so asto bring nature under principles of observation and investigation by analogy : : : withoutpresuming to explain it : : : (ibid.). To take teleology as explanatory would introduce anew causality into natural science, even though in fact we only borrow this causality fromourselves : : : (ibid., p. 237 [5:361]). This would be a quite special kind of causality, or atleast a quite distinct lawfulness [Gesetzmaigkeit] of nature and even experience cannotprove that there actually are such purposes [die Wirklichkeit derselben: : :beweisen] (ibid.,p. 236 [5:359]).

    I think we must be quite cautious in presuming we understand what GesetzmaigkeitandWirklichkeitmeant for Kant, here. Should Gesetzmaigkeit be translated as lawfulnessor lawlikeness? Are these dierent for Kant, as similar compounds with Zweck have sug-gested to interpreters, that is, can there be lawlikeness in the absence of law? (See Gins-borg, 1997; Rang, 1998; Fricke, 1990; Warnke, 1992). And Wirklichkeit: what is Kantafter in questioning whether we can prove such an actuality? The notion of proving[beweisen] from experience is a very touchy question in transcendental philosophy, butis Kant going even further, suggesting that we can never really be sure that we have expe-rienced a natural purpose, that is, that it is actual? Kant triggers a mineeld of problemsfrom the very moment of phenomenal encounter with organisms. As Marcel Quarfoodexplains, organisms like all objects of experience are subject to the causal principle,but there are features of organisms that appear to be intractable for the kind of explana-tions in terms of causal laws appropriate for ordinary physical objects and thus there isno explanation (or law) for how matter comes together in the ways characteristic fororganisms (Quarfood, 2004, p. 146). Kant characterizes what presents itself as an organ-ism provisionally, [as] a thing [that] is both cause and eect of itself (Kant, 1987, p. 249[5:370]). While we can think this causality without contradiction, we cannot grasp [begrei-fen] it (ibid., [5:371]). That is, we cannot bring it under concepts of the understanding.

    The distinction between erklarbar and erkennbar is of crucial signicance for the loos-ening of the theoretical structure of Kants philosophy by virtue of the Critique of teleo-logical judgment. That one can recognize what one cannot explain puts an enormousstrain on the coherence of experience in Kants system. That teleology can be an expla-nation of phenomenal elements of nature is equally straining. Teleology in its identica-tory role singles out the biological object as a functional unit, an organism : : : Thequasi-explanatory use of teleology serves to provide a law, or at least some order, uni-fying the vast number of causal chains that interact in the organism in ways otherwisewholly contingent (Quarfood, 2004, p. 157). Technically, Kant must deny that teleologycan explain anything in phenomenal nature (compare Simon, 1976; Flasch, 1997; Fricke,1990; Ginsborg, 1987). That is what it means to privilege the mechanistic maxim. Whatteleology is alone permitted to do is oer an analogy with heuristic utility. It is even lessthan an empirical conjecture.

    Many historians and philosophers of biology credit Kant with an elucidation of orga-nized being [organisiertes Wesen] that was enabling for the crystallization of biology as a

    special science (Lieber, 1950; Baumanns, 1965; Low, 1980; Huneman, 2002; Lorenz, 1975;

    LeonardoHighlight

    LeonardoHighlight

    LeonardoHighlight

  • couldor hyoutsiwhatdistinphysiThe pthe organism. When the tree assimilates inorganic into organic matter: it transforms it into

    a part of itself, and hence the tree becomes a product, not merely an educt. Kant recog-nized that metabolism is a matter of systemic (re)integration. He distinguished emphati-cally between growth by internal integration and mere aggregation. Mechanism can getno further than aggregation; but systems do more: The whole is thus an organized unity

    (articnever do of itself, for that would be generatio aequivocaspontaneous generationlozoism). He went on: in terms of the ingredients that the tree receives from naturede it we have to consider it to be only an educt (ibid.). An educt merely redeploysis already given. By contrast, a product makes something new. In terms of Lockesction, all organic parts are compounded of particles that are drawn from the generalcalchemical order and these particles are as such indierent to organic assimilation.oint is the converse: something really does become dierent, but it is at the level ofReill, 2004). But Peter McLaughlin has given a very useful account of the development ofeighteenth-century thought about organism which would give credit for the decisiveinsight rather to John Locke (McLaughlin, 2001, pp. 173178). As McLaughlin construesLocke, he

    attempted to determine the dierence between the identity over time of an aggregateor mass and that of an organism. The identity of a material substance or a phys-ical body consists in the identity of its parts and thus ultimately in the identity of itscomponent atoms: : : The identity of an organism, on the other hand, consists not insubstantial identity but rather in vital (modal) identity, in the identity of the processof continued renewal and regeneration of the parts of the system by the system.(Ibid., pp. 175176; Locke, 1998 [1689])

    McLaughlin continues: Locke here introduces two levels of organization: parts and par-ticles : : : Life consists in the ability of the system to renew (continue and frame) its parts: : : An organic body is now conceptualized as something that remains identical to itself : : :by continually reproducing itself and its parts : : : (McLaughlin, 2001, p. 176). (Thus, thereplacement of one carbon atom by anotherat the particle levelaects nothing regard-ing the identity of the part or the organism.) This was the decisive eighteenth-century re-trieval and reformulation of Aristotles notion of entelechy, of intrinsic purposiveness.Buon took it up into eighteenth-century empirical life science, especially in terms ofhis notion of the moule interieure (Sloan, 1979). Kant inherited it from the ensuing discus-sions among life scientists of the day.

    In the third Critique, Kant, on the example of a tree, highlighted three such featuresthat eighteenth-century life science considered empirically established about organisms.First, with regard to its species the tree produces itself: with its species, it is both causeand eect, both generating itself and being generated by itself ceaselessly (Kant, 1987,p. 249 [5:371]). Here, Kant slid between the type or species tree and its individual tokensin asserting the reciprocity of cause and eect. For our part, what is important is that inpropagation Kant saw a functional property that is intrinsic to organisms.

    Next, he highlighted physiological functions of more properly individual organisms:nutrition and growth. Kant stressed the particular point that the matter that the tree assim-ilates is rst processed by it until the matter has the quality peculiar to the species, a qual-ity that the natural mechanism outside the plant cannot supply (ibid.). That is, the treetransforms inorganic materials into organic forms (something he believed inorganic matter

    J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 757ulatio), and not an aggregate (coacervatio). It may grow from within (per intussuscep-

  • 758 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770tionem), but not by external additions (per appositionem) (Kant, 1923, p. 653 [3:539]). Forthis reason, Kant rejected the notion that crystal growth was proto-organic; it was in hisview strictly mechanical.

    The contrast of educt and product is crucial, as Kant elaborated later in the third Cri-tique and as I have explored in my discussion of his view of epigenesis, which turns on thisdistinction (Kant, 1910, Vol. 5, p. 423; Zammito, 2003; Zammito, 2006). Kant continued:the separation and recombination of this raw material show that these natural beings havea separating and forming ability [Scheidungs- und Bildungsvermogens] of very great origi-nality; all our art nds itself innitely outdistanced : : : (Kant, 1987, p. 250 [5:371]). Thisemphasis on the creative spontaneity of nature as incomparably superior to human articeis repeated in a decisive passage later in the third Critique, which substantially disqualiesthe analogy of design in our conceptualization of organisms (ibid., p. 254 [5:374375]).

    The last of the distinguishing traits of organisms Kant formulated on the model of thetree was the power of regeneration, that is, healing and even organ restorationa matterof the most exciting experimental discoveries of mid-eighteenth-century empirical life sci-ence: the experiments of Trembley, Bonnet, and ultimately Blumenbach (Vartanian, 1950;Roger, 1963 and Roe, 1981). What the regenerative powers of the polyp conrmed werethe arguments that Caspar Friedrich Wol had advanced based on his embryologicalobservations, namely that organic forms had the power of developmental self-organiza-tion, a capacity for internal regulation and modulation which involved a continuousand mutually responsive relation between parts and whole for the sustenance of the organ-ism. This was the essential argument of eighteenth-century epigenesis.

    These features of intrinsic purposiveness Kant discerned in organisms seem distinctlyphysiological, that is they have to do more with questions of development or maintenancethan of origination. To phrase it in the terms of causal role theory, what Kant highlightsin organic systems is persistence and plasticity in securing system-maintenance. As Steiger-wald construes Kant, he highlights their exibility in the realization of [their] forms andfunctions, the degree of freedom or at least spontaneity in their operations, which nev-ertheless demonstrate regularities in formation, structure and functioning (Steigerwald,2006, this issue).

    These marvelous properties of organized creatures are part of the empiricalexperien-tial data available to human investigators trying to comprehend the order of nature. Thatis, Kant appears to consider their phenomenal description unproblematic. But how thesemarvelous properties can be explainedand how they can be integrated into a systemof empirical laws as the order of natureremains, for Kant, a philosophical conundrum.That an entity can be cause and eect of itself, Kant argued, is beyond discursive rational-ity. Yet that is what is required to conceive a natural purpose. Steigerwald writes: In orderfor us to judge a body as a natural purpose, not only is it necessary that we conceive thepossibility of the parts as dependent for their existence and their form on their relation tothe whole, but also that all the parts through their own causality reciprocally produce oneanother as regards their form and combination and in this way produce a whole (Steiger-wald, 2006, this issue). To be sure, Kant articulated this point, but he did so in order to setit beyond human explanatory power.

    Quite simply, for an entity to be cause and eect of itself is for it to become inexplicableby all the resources of science according to the nexus eectivus. But what, exactly, is thisnexus eectivus? If we wish to understand what Kant could possibly have meant in claiming

    that we can at one and the same time experience organisms and yet not be able to explain

    LeonardoHighlight

    LeonardoHighlight

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 759them within the order of natural law, we must drive a wedge between experience and expla-nation, that is, not simply between description and explanation as a posteriori judgments,but between the a priori binding character of determinant judgment for an object in generalto be part of a nature possible for us to experience, and the multiply contingent character ofreective judgment, that is, a posteriori scientic explanation.Many of us have been drawnto think that Kant drastically compromised the force of his Second analogy account ofcausality (as determinant or constitutive) by his representation of mechanical causality inthe third Critique as merely reective or regulative. We were misguided to think that Kantalways meant the same thing by mechanism, or indeed, by the causal nexus.

    A persuasive account is now in place that Kant intended a more restricted sense ofmechanical causality in the third Critique. McLaughlin has argued that Kant came toa new realization that mechanical causality as practiced by natural science introduced asupplementary restriction whereby the causal relation between parts and wholes couldonly work aggregatively from parts to whole, and not mutually, or inversely (McLaughin,1990). Quarfood summarizes McLaughlins claim elegantly: whereas constitutive causal-ity involves a determination of time-order, the maxim of mechanism is a further, regulativeassumption about the relation of part to whole, according to which a whole is always caus-ally determined by its parts (Quarfood, 2004, p. 161). In a very inuential essay, Allison(1991) has seconded this interpretation. Not everyone agrees with McLaughlins specicaccount (the partwhole argument). Hannah Ginsborg has argued that this is insucientto distinguish organisms from machines, or crystals from organisms. She resorts to notionsof normativity to complicate the causal theory (Ginsborg, 2001, 2004). But even Ginsborg,and with her others like Quarfood (2004), Steigerwald (this volume), and Breitenbach (thisvolume), construe Kant as employing a dierent sense of mechanism in the third Critique,so that now this is the reigning wisdom. Causality, at the determinant or constitutive level,remained fully general for Kant, just as he contended in the rst Critique, and so it waspossible to experience this bizarre phenomenon of something as cause and eect ofitselfthough only by immediately translating it into an analogy that was prima facieinadequate. But scientic explanation could not accommodate it, insofar as it restricteditself to the mechanistic maxim.

    Now, clearly Kant committed himself without reservation to the appropriateness of fol-lowing this maxim of mechanical explanation. He only introduced the epistemologicalproviso that, as niteindeed, discursiveintelligences, we might never be able to bringit to closure. The unity of science as an order of nature in its empirical laws constituted aregulative ideal we needed to believe in to pursue science, but all the same it might never befullled. One can take this as a pretty fair anticipation of the underdetermination thesisespecially in Quines strongest form (Quine, 1975). What is striking is that, precisely on thebasis of organisms, Kant felt so condent that he claimed (what Quine was forced to repu-diate) that it was already inescapable that the project must fail. Quarfood formulates theparadox (or could it be a contradiction?): how Kant can claim both that organisms aremechanically unexplainable and that it nonetheless is possible that they are mechanicallyproduced (Quarfood, 2004, p. 196). If there is an antinomy of judgment it is about thismechanist perplexity far more than about teleology per se (McLaughlin, 1990; Allison,1991; Philonenko, 1977; Zanetti, 1993). Can there be an acceptable order of nature ifwe have to exclude something as pervasively present as life forms?

    What does this experience incapable of explanationor only of a knowingly falla-

    cious explanation (analogy of design)actually refer to? What is it about? Is the organism

  • 760 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770a fact or a confusion of experience? The very discernment of an order, of holism, seems todemand something far more than an analogy to artice, as Kant himself acknowledged.As Ginsborg puts it, the question remains of how we can even coherently regard some-thing both as a purpose and as natural (Ginsborg, 2001, p. 236). The appeal to analogydoes not overcome the diculty, she continues (ibid., p. 238). Kant himself admitted it:Strictly speaking, : : : the organization of nature has nothing analogous to any causalityknown to us, that is, intrinsic natural perfection, as possessed by those things that are pos-sible only as natural purposes and that are hence called organized beings, is not conceivableor explicable on any analogy to any known physical ability, that is, ability of nature, notevensince we belong to nature in the broadest senseon a precisely tting analogy tohuman art (Kant, 1987, p. 254 [5:375]).

    Perfection as a conceptualization of intrinsic purposiveness opens up the dimension ofnormativity that has become so urgent in contemporary controversy. Ginsborg highlightsKants language in 68 of the third Critique: natural purposes are singly and alone explica-ble according to natural laws which we can think to ourselves only under the idea of purposeas a principle, and indeed merely in this way are they so much as internally cognizable asregards their inner form (Ginsborg, 2001, p. 233; Kant, 1910, Vol. 5, p. 383). Kant arguesin the sentence before that physics needs to avoid crossing the boundary into metaphysics.But does his sentence not cross that boundary, and perhaps more than once? My concernsare (1) what itmeans within the critical philosophy to conceive of natural laws in this fash-ion; (2) what it means that something be internally cognizable as regards inner form.What does this double inwardness betoken? Perfection in its eighteenth-century Germanscholasticphilosophical (and thus, Kants) sense holds the clue. Ginsborg redeems the ideaof natural purpose by insisting that to regard something as a purpose is to regard it as sub-ject to normative rules or standards. She characterizes this kind of lawfulness (or lawlike-ness) in natural purpose as normative law, as distinguished from the necessity either ofconstitutive cognitive principles or of moral obligation (Ginsborg, 2001, p. 232). Kant him-self introduced the notion of what something was supposed to be in a key passage of hisFirst Introduction to the Critique of judgment (Kant, 1910, Vol. 20, p. 240). That appearsto signify that purpose cannot be restricted simply to the concept as cause of the existence ofsome object, as Kant formulated it in 10 of the third Critique, but that it characterizessomething intrinsic to the object itself. The force of Ginsborgs argument suggests we musttake the normativity to be immanent in the object, a product of nature, not simply a rule ofdesign [Bauplan] analogically referred to a nonexistent designer, as in Kants metaphor ofa Technik der Natur. Like contemporary systems theory or theory of self-organization,Ginsborg seems to ascribe normativity to the actual and the empirical order, notwithstand-ing Kants notorious insistence that this can only be as if.

    Why is this normativesystemic immanence so intractably unknowable for Kant, andwhy does he believe that analogy to design is our onlyalbeit inadequatealternative?(see Roque, 1985; Gfeller, 1998; Krohn & Kuppers, 1992). To what is the analogy reallybeing made? It is not the work of art, but the artist (human agency). Kant suggested fornatural purpose a remote analogy with our causality in accordance with ends (Kant,1987, p. 255 [5:375]). But is the analogy really remote in Kants argument? More point-edly, is our causality transparent to our awareness? What do we really know about that?Man is conscious of himself as a self-moving machine, without being able to furtherunderstand such a possibility Kant wrote in his Opus postumum (Kant, 1910, Vol. 21,

    p. 213). How did Kant understand it?

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 761Kant developed an elaborate concept of systemic integration, for which organism wasbut one instance (Zammito, 1992; Guyer, 1990, 2000, 2001, 2003). His primary instancewas reason itself. But this allowed Kant to draw a fundamental analogy between organismand reason as systems (Kant, 1910, Vol. 3, p. 538539). Quarfood recognizes the impor-tance to Kant of the biological analogy : : : between the systematicity of reason and thefunctional integration of the parts in an organism (Quarfood, 2004, p. 79). Kants essen-tial transcendental claim is that reasons principles are a priori, that they cannot be derivedempirically. This was the thrust of his famous analogy between pure reason and epigenesisat B 167 of the rst Critique (Kant, 1910, Vol. 3, p. 128; Haner, 1997; Ingensiep, 1994;Duchesneau, 2000). Quarfood insists that the comparison of transcendental philosophyand biological epigenesis is [strictly] an analogy, that it would be wrongheaded to thinkthat one could make any determinate inference from biological science to transcendentalphilosophy, as he (mistakenly) accuses Phillip Sloan to have done (Quarfood, 2004, pp.79112; Sloan, 2002). The question, from our vantage, Quarfood notwithstanding, is:can this matter be left at the level of mere analogy? Logically, analogy postulates at leastsome common concept under which two quite dierent matters can be subsumed, somecorrespondence at least in terms of a relation. But are reason and organisms simply speciesof some logically higher genus [Schulgattung] of system? Or are they species of a generativeorder [Naturgattung], such that reason is an expression of the particular organism calledman? That, of course, is the gist of the naturalist program in the philosophy of mind.

    How should we understand reason in Kant? What is its nature? Whence comes this (epi-stemic) power [Vermogen or Kraft]? What makes it real? Has Quarfood achieved anymetaphysical cost saving (or Kant before him) in conceiving it as an acquisitio origina-ria? (Quarfood, 2004, pp. 77117). What, and more pertinently, how is that? Here, Kantsanalogy of organism and reason requires further inquiry. Guyer writes: it is only ourawareness of the freedom of our own purposiveness that leads us to conceive of the pur-posiveness of organisms as necessitating a fundamental split between the teleological andmechanical views of nature (Guyer, 2001, p. 264). And he then asks the very pertinentquestion: Is it just an empirical fact, a matter of empirical psychology rather than tran-scendental psychology, that organisms suggest the idea of purposiveness to us? (ibid.,p. 276).

    Consider the conative conception of reason which scholars have begun to discern inKants writings (Yovel, 1989; Gilead, 1985; Doringer, 2000). Kleingeld (1998) has oereda very careful consideration of Kants strange usage of needs or interests of reason andfurther talk of the right of such impulses. What is to be made of that? Kleingeld identiesthree possible senses. First, might it be an anthropological recasting of Kants transcen-dentalismthat is, is reason merely human? Second, could all this be simply a matterof metaphorical language, warranting no further philosophical concern? (See Nuyen,1989). Finally, might this be more of a root metaphor, that is, might it betoken the resur-facing of metaphysical elements in Kants transcendentalism? Is Kants supersensibleground here demonstrating agential properties, though not necessarily human ones (inHegelian terms, must we render reason as Geist)? Kleingeld holds that none of theseoptions is satisfactory. To account for Kants language she resorts to two moves. First,in assessing Kants talk of the right to recognize needs or interests of reason, she arguesthat he is clear that such subjective impositions are permissible only when there are noobjective determinations governing a matter. Thus, they arise primarily with reference to

    the supersensible, and take on a practical-rational register: orientation in thinking for

  • 762 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770the sake of practical obligations (Kant, 1910, Vol. 8, pp. 131148). Further, Kleingeldargues that Kant oers us a philosophical basis for interpreting his usage of need andinterestnamely his theory of symbolic hypotyposis (ibid. Vol. 5, p. 374). Ideas of reasoncannot be fully instantiated in intuition, but they can be evoked analogically in a symbol.Kleingeld argues that reason is itself an idea of reason in Kants technical sense: we can-not formulate a determinant judgment about it because it is not present to intuition.

    There is no question that Kant wanted to eschew metaphysics, with reference to nat-ural purposes as with reference to many other concepts; the question is whether hisown system allowed that. I suggest there is a passage in the Critique of teleological judg-ment that brings all this to a decisive head: In considering nature and the ability it dis-plays in organized products, we say far too little if we call this an analogue of art, forin that case we think of an artist (a rational being) apart from nature. : : : We might becloser if we call this inscrutable property of nature an analogue of life, Kant wrote(1987, p. 254 [5:374]). This seems to a modern reader bizarre: life is what we think organ-ism is already about, so what analogy could there be? Kant continues: But in that case wemust either endow matter, as mere matter, with a [kind of] property (hylozoism) that con-icts with its nature. Or else we must supplement matter with an alien principle (a soul)conjoined to it (ibid.). Life, it appears, comes with heavy metaphysical baggage. Whatis the proper scope of Leben and what is its philosophical meaning (that is, what is theextension and what is the intension of the idea)? Kant is adamant that brute matter cannotpossess this character. The essence of matter is inertia: all change in motion must have anexternal cause. To ascribe to brute matter the inner capacity to inaugurate motion wouldbe the death of natural philosophy (Kant, 1910, Vol. 4, p. 544). Hylozoism was anath-ema to Kant. But what alternative did the analogy of life oer him for explaining organ-ism? For Kant, brute matter was soulless and dead; man was a living organism. Whatabout the intermediate orders: what of animals, what of plants? The point is, if the conceptof life is restricted to intelligent will, is man the only living thing? (Ingensiep, 2004, p.121).

    One must be cautious in arming that Kant was not introducing a special matter orforce (Steigerwald, 2006, this issue). In his metaphysics lectures, Kant was more explicitthan he permitted himself to be in the third Critique: all matter that is animate has aninner principle which is separated from the object of outer sense, and is an object of innersense : : : Thus, all matter which lives is alive not as matter but rather has a principle of lifeand is animated. But to the extent matter is animated, to that extent it is ensouled (Kant,1910, Vol. 28, p. 275). Ingensiep (2004) argues that Kants careful arguments in the Para-logisms of the rst Critique banned the notion of an immortal soul, but not that of anembodied one. If the human organism was endowed with life, this was because its matterwas as ensouled as its soul was embodied. What Aristotle identied with soul, Kanthad to reformulate to bet a modern science context. Late in his life, in Perpetual peace,Kant acknowledged the popularity of the new term Lebenskraft, which he recognized as asubstitute for the traditional concept of the soul. He welcomed this term-shift, for he heldthat while we can recognize an eect, we should be wary of hypostatizing a substance as itsground (Kant, 1910, Vol. 8, p. 413).

    Ingensiep suggests that we array Kants many sayings about life in terms of the Aristo-telian scale of being, from brute matter through plants and animals to man and to theworld as a whole. Of the many statements Kant made about life, a good starting point

    is from his Critique of practical reason: Life is the power of a being to act in accordance

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 763with the laws of the faculty of desire (Kant, 1956, p. 9 n. [5:9 n.]). The faculty of desire, inturn, is the power to be through its representations the cause of the actuality of these rep-resentations (ibid.). No one can miss the parallel between this denition of the faculty ofdesire and Kants denition, in the third Critique, of the key term purpose. Thus the ques-tion of the actuality (or actualization) of the object through a purpose or the faculty ofdesire entails a theoretical component. In his early Dreams of a spirit-seer, Kant wrote:all life consists in the inner capacity of self-determination according to free choice [Will-kur] (ibid, Vol. 2, p. 327). In his Reections Kant observed: life is nothing but faculty ofdesire in its minimal exertion [in der geringsten Ausubung] (ibid. Vol. 15, p. 465). In hisOpus postumum, Kant wrote: life in the strictest meaning of the term is the capacity ofsponaneity of a physical entity to act in accordance with certain of its own representations(ibid. Vol. 20, p. 566). Kant seemed to be willing to extend at least some measure ofdesireaction in accordance with representationsto animals, though it is desire entirelydriven by pleasure/pain, and not by rational choice. Even were we to grant that, what ofplants? As Ingensiep stresses, Kant never ascribed psychological desire, even analogically,to plants.

    Thus, plants epitomize Kants conceptual discrimination of life from organism. As wehave seen, he illustrated the features of organism precisely by a planta tree. Conse-quently, Kants notion of organism is broader than that of life, and the failure of thesetwo terms to have the same extension expresses the insuciency Kant acknowledged inhis analogy of life for natural purpose. How do we construe the residual disanalogyfor biology? Blumenbachs notion of Bildungstrieb won Kants endorsement for all organicforms. Animals were clearly identied with Trieb, but even plants have a Bildungstrieb, notjust Bildungskraft, in the discrimination Kant adopted from Blumenbach (Kant, 1910,Vol. 5, p. 424; McLaughlin, 1982 and Richards, 2000). That is, they have some internal,quasi-spontaneous principle of motion (Ingensiep, 2004, p. 128). The question of Trieb inKants notion of organism denotes the element unaccounted for even by Kants analogy oflife. But what was a Trieb for Kant, and how did he distinguish it from a Kraft? How couldan inner force be actual for scientic inquiry? Kant insisted it could only be heuristic, orregulative. Blumenbach and his school took the Bildungstrieb for actual, not speculative.Their project was to specify its eects through the mechanisms (Bildungskrafte) it set inmotion. Kants regulative/constitutive distinction proved useless for them in that pursuit,though it gave them some metaphysical comfort, especially in the analogy to the Newto-nian mysteriousness of gravity.

    With Robert Butts, I believe it is necessary to hearken to the anomalies that pile up inKants philosophy, and especially his philosophy of science, all of which revolve, as Buttsnotes, around design (Butts, 1990, p. 3). Kant suggested that we could nd our waythrough the diculties by relying on our awareness of our own agential performance.But there are grave questions about whether Kant could warrant this transparency ofhuman agency (theoretical or practical) on his own terms (ibid., p. 15 n. 4). In naturalistterms, to claim we can grasp organism by analogy to our agency is to put the cart beforethe horse, for that agency is contingent upon the features of organism for its own cogency(McLaughlin, 2001, p. 15).

    Organism is a capital anomaly. It will not t in Kants system of science, and yet with-out a good account of it, the system itself must in the end appear inadequate (Zammito,2003). Ingensiep observes: Not only the Kantian but also the modern struggle to discrim-

    inate a conceptual dierence between organism and life leads ultimately back to the

  • roots of entelechy in Aristotelian substance theory (Ingensiep, 2004, p. 136). CanMcLaughlins discrimination of formal from nal cause, and with it the recourse to Aris-totle, provide assistance especially given that the question of the actuality of intrinsic pur-poses in nature (or for science) bears on humansnot merely as knowers but as agents?How did Kant construe intrinsic purposiveness? Did he oer us any conceptual advanceover Aristotle or Locke? To discriminate between form and existence in Steigerwaldsterms, would it not be more fruitful to resort to the Aristotelian distinction between for-mal and material cause, or to Lockes distinction between parts and particles? Theform of the part is a property of a higher level than the material existence of particles thattransiently compose it. The persistence in identity that life science should meaningful expli-cate has to do with parts and organisms, not atoms or elements. Thus one could distin-guish between an organelle and an element, and the questions of identity (andcausality) would be quite dierent. While this still leaves us with circular or holistic cau-sality, it does so at the level of parts, not particles.

    3. Conclusion: philosophy and science: who is the Underlaborer in naturalism?

    Paul Guyer has written that the argument in the third Critique is unstable (Guyer,2001, p. 275). How do we understand that instability? If it was not already post-critical,to use a phrase from Allison (2000), did the third Critique instantiate that notorious gapin the critical system (Forster, 1987) that became Kants preoccupation for the bulk of theperiod after the publication of his rst Critique and the warrant for all subsequent GermanIdealism? What of the Opus postumum, especially according to the aggressive line of inter-pretation advanced by Tuschling (1991) and Edwards (2000)? What of biological practicein Blumenbachs Gottingen school and elsewhere in turn-of-the-century German com-parative morphology, culminating in von Baer and Cuvier? (Richards, 2002; Reill,2004) And, nally, of course, what of Naturphilosophie, that bete noire of positivism?(Cunningham & Jardine, 1990; Gloy, 1994; Gloy & Burger, 1993; Beiser, 2002 and Steiger-wald, 2003) Is it plausible, as some Kantians desire, to privilege the coherence of a partic-ular reading of the rst Critique with a particular reading of the third if it puts all theseother questions beyond the pale?

    It is all well and good to recognize that Kant is talking about the judgments biologistsmake, not the content of their judgments (Ginsborg, 2001, p. 235). But biologists cannot beindierent to the content of their judgments. Quarfood recognizes this clearly. In his dis-sertation he insists upon a distinction : : : between the point of view of the biologist andthe meta level perspective of philosophy (Quarfood, 2004, p. 119). What is striking is notonly that he believes that if a teleological point of view is a condition for biology, teleol-ogy should rather be seen as a constitutive principle for this science (which makes Quar-food a good naturalist), but he thinks Kant held this view! (ibid.) Quarfood asserts that onhis reading of Kant, teleology, though regulative, can be said to be constitutive for biol-ogy, as a condition for the possibility of biological objects (organisms) (ibid., p. 153). Heelaborates: Aristotelian functions : : : are to be seen as valid at the object level of biologicalinvestigation. The denial of their objectivity only takes place at a further meta level reec-tion. For the biologist, organisms exhibit functions as a matter of course (ibid., p. 152). Ican nd no passage in Kant that warrants ascribing to him this concession to biologicalpractitioners. On the contrary: Kant saw himself called to caution them about their

    764 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770daring adventure of reason (Kant, 1910, Vol. 5, p. 419 n.). The charm is that the biol-

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 765ogists misunderstood his admonitions in a way that enabled them to get on with theirwork.

    Kantdrawing on his eighteenth-century predecessorsprovided a discerning andpowerful characterization of what biologists had to explain in organic form. His dierencefrom the rest is that he opined that it was impossible to explain it. All one could do wasdraw poor analogies to it. Its inscrutability was intrinsic. That is the implication of ClarkZumbachs provocative title: The transcendent science (Zumbach, 1984). We need to takeKants assertion of inscrutability seriously. Kant set a hard and fast boundary markerbetween attainable science and speculative metaphysics. But did he mark the boundaryproperly? Need we halt there? Have we halted there? Kant found this boundary very usefulin pursuing his larger philosophical project. I submit that for biologists it was hardly sopropitious, even if they really tried to work with it (Zammito, 1998). Only by misunder-standing Kant did biology as a special science emerge at the close of the eighteenth cen-tury. Robert Richards put it succinctly: The impact of Kants Kritik der Urteilskraft onthe disciplines of biology has, I believe, been radically misunderstood by many contempo-rary historians. : : : Those biologists who found something congenial in Kants third Cri-tique either misunderstood his project (Blumenbach and Goethe) or reconstructed certainideas to have very dierent consequences from those Kant originally intended (Kielmeyerand Schelling) (Richards, 2002, p. 229). And my judgment is that for philosophers of biol-ogy today, especially naturalists, Kant has little to oer. Locke and the other eighteenth-century theorists of organisms well may, but not Kant.

    For Locke and a fortiori for Buon and his followers intrinsic purposiveness was a factof the matter about concrete biological phenomena; the features of internal self-regulationwere hypotheses arising out of actual research practice. There is, to be sure, a blurredboundary between where hypothesis ends and where speculation beginsand anxietiesabout that boundary were as acute in an eighteenth-century Newtonian context wherefeigning no hypotheses was sacrosanct, as in a (post-)positivist context at the close ofthe twentieth century leery of metaphysical costs. But scienceas Buon alreadyinsistedis probabilistic, not apodictic (Sloan, 1985, 1995). It is about inference to the bestexplanation, always as contingent as it is fallible, but not for all that either arbitrary orhopeless.

    But there is another feature that needs to be stressed, namely the relation of so-calledspecial sciences to physics. The unity of science has now been vigorously challenged as ahistorical and cultural ction. It is the disunity of science we nd whenever we take upserious investigation of the practices of science (Galison & Stump, 1996). And that sug-gests that special sciences need feel no longer the positivist shudder of submission tohegemonic physical reduction. Naturalism is obliged to nd a way to conceptualize a spe-cial science whose terms exceed physicalmechanical reductionism and yet achieve a deter-minacy of empirical lawfulness that deserves accreditation (a notion of special science inwhich biology might be simply one of a number of cases). I suggest that we take seriouslyenough the type/token distinction to consider a new conceptualization of the organizationof the various sciences. Here I follow some very promising proposals from Tucker (2004,p. 260), who aligns the new evolutionary biology with a larger group of sciences (includinghistory) which take as their object of inquiry tokens, not just types, and therefore ndthemselves inevitably caught up in questions of historicism.

    In the epigraph above, McLaughlin makes the claim that sciences ought to dene their

    own appropriate projects and practices, and that it is the role of philosophy simply to

    LeonardoHighlight

  • dered foundationalismeven Kantsand more emancipated, for our collective enter-

    766 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770prise is from the outset only provisional, but it should provision us as boldly as ourresources permit. I submit a philosopher of science who pronounces from the epistemolog-ical high ground has no place with us in Neuraths boat on the naturalist sea.

    References

    Adickes, E. (1924). Kant als Naturforscher. Berlin: de Gruyter.Allen, C., Beko, M., & Lauder, G. (Eds.). (1998). Natures purposes: Analyses of function and design in biology.

    Cambridge, MA: MIT Press.Allison, H. (1991). Kants antinomy of teleological judgment. Southern Journal of Philosophy, 30 (Suppl.), 2542.Allison, H. (2000). Is the Critique of judgment post-critical? In S. Sedgwick (Ed.), The reception of Kants critical

    philosophy (pp. 7892). Cambridge: Cambridge University Press.Baumanns, P. (1965). Das Problem der organischen Zweckmaigkeit. Bonn: Bouvier.Bedau, M. (1991). Can biological teleology be naturalized? Journal of Philosophy, 88, 647655.Bedau, M. (1992). Goal-directed systems and the good. The Monist, 75, 3451.Bedau, M. (1993). Naturalism and teleology. In S. Wagner, & R. Warner (Eds.), Naturalism: A critical appraisal

    (pp. 2351). Notre Dame: University of Notre Dame Press.Beiser, F. (2002). German idealism: The struggle against subjectivism, 17811801. Cambridge, MA & London:

    Harvard University Press.Bigelow, J., & Pargetter, R. (1998). Functions. In C. Allen, M. Beko, & G. Lauder (Eds.), Natures purposes:

    Analyses of function and design in biology (pp. 241260). Cambridge, MA: MIT Press.Buller, D. J. (Ed.). (1999). Function, selection and design. Albany: SUNY Press.Butts, R. (1990). Teleology and scientic method in Kants Critique of judgment. Nous, 24, 116.Christensen, W. (1996). A complex systems theory of teleology. Biology and Philosophy, 11, 301320.Churchland, P. (1995). The engine of reason, the seat of the soul: A philosophical journey into the brain. Cambridge,

    MA: MIT Press.Cummins, R. (1975). Functional analysis. Journal of Philosophy, 72, 741764.Cunningham, A., & Jardine, N. (Eds.). (1990). Romanticism and the sciences. Cambridge: Cambridge Universityassess the metaphysical cost. I think that comes close to the core principle of naturalismin philosophy of science. It bespeaks more authentically than logical positivism ever didor, I am suggesting, Kant eitherthe proper relation between philosophy and science inthe modern world. It is not surprising that Kants epistemological discretion should appealto philosophers of science today, especially as the conventional wisdoms of positivism per-ish in the deep, dogma by dogma. However, that it should ultimately satisfy current prac-titioners of biological science is, I suspect, just as problematic now as it proved then.Perhaps the same misunderstanding of what Kant restricted to the merely reectivemight prot biologyepistemological skeptics notwithstandingeven today. If biologymust conceptualize self-organization as actual in the world, Kants regulative/constitutivedistinction is pointless in practice and the (naturalist) philosophy of biology has urgentwork to undertake for which Kant turns out not to be very helpful.

    Kant is harbor only for those who seek epistemological shelter from the hard problemsof function talk. Once we slip Kants safe epistemological moorings, we enter into trou-bled waters, but they are our waters: these are the questions of our naturalism. Where the-ory stands between epistemology and metaphysics is a very dicult post-positivistquestion: both theory-ladenness of observation and underdetermination of theory rela-tive to available or possible evidence bear directly upon it. Naturalist philosophers of biol-ogy today need not succumb to the scruples (whether absolute or transcendental) thathaunted eighteenth-century philosophers. We are both more scrupulous in having surren-Press.

    LeonardoHighlight

    LeonardoHighlight

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 767Depew, D., & Weber, B. (Eds.). (1985). Evolution at the crossroads: The new biology and the new philosophy ofscience. Cambridge, MA: MIT Press.

    Doringer, B. (2000). Das Leben theoretischer Vernunft. Berlin: de Gruyter.Duchesneau, F. (2000). Epigene`se de la raison pure et analogies biologiques. In F. Duchesneau, G. Lafrance, &

    C. Piche (Eds.), Kant actuel (pp. 233256). Montreal: Bellarmin.Edwards, J. (2000). Spinozism, freedom, and transcendental dynamics in Kants nal system of transcendental

    idealism. In S. Sedgwick (Ed.), The reception of Kants critical philosophy (pp. 5477). Cambridge: CambridgeUniversity Press.

    Flasch, W. (1997). Kants Empiriologie: Naturteleologie als Wissenschaftstheorie. In P. Schmid, & S. Zurbuchen(Eds.), Grenzen der kritischen Vernunft (pp. 273289). Berlin: Schwabe.

    Forster, E. (1987). Is there a gap in Kants critical system? Journal of the History of Philosophy, 25, 533555.

    Fricke, C. (1990). Explaining the inexplicable: The hypotheses of the faculty of reective judgment in Kants thirdCritique. Nous, 24, 4562.

    Galison, P., & Stump, D. (Eds.). (1996). The disunity of science: Boundaries, contexts, and power. Stanford:Stanford University Press.

    Genova, A. (1975). The purposive unity of Kants critical idealism. Idealistic Studies, 5, 177190.Gfeller, T. (1998). Wie tragfahig ist der teleologischen Bruckenschlag? Zu Kants Kritik der teleologischen

    Urteilskraft. Zeitschrift fur philosophische Forschung, 52, 215236.Gilead, A. (1985). Restless and impelling reason: On the architectonic of human reason according to Kant.

    Idealistic Studies, 15, 137150.Ginsborg, H. (1987). Kant on aesthetic and biological purposiveness. In A. Reath, B. Herman, & C. Korsgaard

    (Eds.), Reclaiming the history of ethics (pp. 329360). Cambridge: Cambridge University Press.Ginsborg, H. (1997). Lawfulness without a law: Kant on the free play of imagination and understanding.

    Philosophical Topics, 25, 3781.Ginsborg, H. (2001). Kant on understanding organisms. In E. Watkins (Ed.), Kant and the sciences (pp. 231258).

    Oxford: Oxford University Press.Ginsborg, H. (2004). Two kinds of mechanical inexplicability in Kant and Aristotle. Journal of the History of

    Philosophy, 42, 3365.Gloy, K. (1994). Die Naturauassung bei Kant, Fichte und Schelling. FichteStudien, 6, 253275.Gloy, K., & Burger, P. (Eds.). (1993). Die Naturphilosophie im Deutschen Idealismus. Bad Cannstatt: Frommann-

    Holzboog.Godfrey-Smith, P. (1993). Functions: A consensus without unity. Pacic Philosophical Quarterly, 74, 196208.Godfrey-Smith, P. (1994). A modern history theory of functions. Nous, 28, 344362.Gould, S. J., & Lewontin, R. (1979). The spandrels of San Marco and the Panglossian paradigm: A critique of the

    adaptationist program. Proceedings of the Royal Society, B205, 581598.Gould, S. J., & Vrba, E. (1982). Exaptationa missing term in the science of form. Paleobiology, 8, 415.Griths, P. (1996). The historical turn in the study of adaptation. British Journal for the Philosophy of Science, 47,

    511532.Griths, P., & Gray, R. (1994). Developmental systems and evolutionary explanation. Journal of Philosophy, 91,

    277304.Guyer, P. (1990). Reason and reective judgment: Kant on the signicance of systematicity. Nous, 24, 1743.Guyer, P. (2000). The unity of nature and freedom: Kants conception of the system of philosophy. In S.

    Sedgwick (Ed.), The reception of Kants critical philosophy (pp. 1953). Cambridge: Cambridge UniversityPress.

    Guyer, P. (2001). Organism and the unity of science. In E. Watkins (Ed.), Kant and the sciences (pp. 259281).Oxford: Oxford University Press.

    Guyer, P. (2003). Kant and the systematicity of nature: Two puzzles. History of Philosophy Quarterly, 20,277295.

    Haner, T. (1997). Die Epigenesisanalogie in Kants Kritik der reinen Vernunft. Ph.D. thesis, Universitat desSaarlands.

    Huneman, P. (2002). Metaphysique et biologie: Kant et la constitution du concept dorganisme. Villeneuve: PressesUniversitaires du Septentrion.

    Ingensiep, H. W. (1994). Die biologische Analogien und die erkenntnistheoretische Alternativen in Kants Kritikder reinen Vernunft B 27. Kant-Studien, 85, 381393.

  • 768 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770Ingensiep, H. W. (1998). Organismus und Leben bei Kant. In H. W. Ingensiep, H. Baranzke, & A. Eusterschulte(Eds.), Kant-Reader: Was kann ich wissen? Was soll ich tun? Was darf ich hoen? (pp. 107136). Wurzburg:Konigshausen & Neumann.

    Kant, I. (1910). Kants gesammelte Schriften (Koniglich Preussischen Akademie der Wissenschaften, Ed.) (29vols.). Berlin: Reimer.

    Kant, I. (1923). Critique of pure reason (N. Kemp-Smith, Trans.). Toronto: Macmillan, 1923.Kant, I. (1956). Critique of practical reason (L.W. Beck, Trans.). Indianapolis: Bobbs-Merrill.Kant, I. (1987). Critique of judgement (W. Pluhar, Trans.). Indianapolis: Itackett.Kleingeld, P. (1998). The conative character of reason in Kants philosophy. Journal of the History of Philosophy,

    36, 7797.Krohn, W., & Kuppers, G. (1992). Die Naturlichen Ursachen der Zwecke: Kants Ansatze zu einer Theorie der

    Selbstorganisation. Selbstorganisation: Jahrbuch fur Komplexitat in der Natur-, Sozial-, und Geisteswissens-chaften, 3, 3150.

    Lagier, R. (2004). Les races humaines selon Kant. Paris: PUF.Lewontin, R. (1985). The organism as subject and object of evolution. In R. Levins, & R. Lewontin (Eds.), The

    dialectical biologist (pp. 85106). Cambridge, MA: Harvard University Press.Lewens, T. (2001). No end to function talk in biology. Studies in History and Philosophy of Biological and the

    Biomedical Sciences, 32, 179190.Lewens, T. (2004). Organisms and artifacts: Design in nature and elsewhere. Cambridge, MA: MIT Press.Lewens, T. (Forthcoming). Functions. In M. Matthen, & C. Stephens (Eds.), Handbook of the Philosophy of

    Science: 3: Philosophy of Biology. Amsterdam: Elsevier.Lieber, H.-J. (1950). Kants Philosophie des Organischen und die Biologie seiner Zeit. Philosophia Naturalis, 1,

    553570.Locke, J. (1998). Essay concerning human understanding. Harmonsworth: Penguin (First published 1689).Lorenz, K. (1975). Kants doctrine of the a priori in the light of contemporary biology. In R. Evans (Ed.), Konrad

    Lorenz: The man and his ideas (pp. 181217). New York & London: Harcourt Brace Jovanovich.Low, R. (1980). Philosophie des Lebendigen: der Begri des Organischen bei Kant, sein Grund und seine Aktualitat.

    Frankfurt: Suhrkamp.McLaughlin, P. (1982). Blumenbach und der Bildungstrieb: Zur Verhaltnis von epigenetischen Embryologie und

    typologischen Artbegri. Medizinhistorisches Journal, 17, 357372.McLaughin, P. (1990). Kants critique of teleology in biological explanation: Antinomy and teleology. Lewiston:

    Mellen.McLaughlin, P. (2001). What functions explain: Functional explanation and self-reproducing systems. Cambridge:

    Cambridge University Press.Millikan, R. (1989). In defense of proper functions. Philosophy of Science, 56, 288302.Millikan, R. (1997). White Queen psychology and other essays for Alice. Cambridge, MA: MIT Press.Nagel, E. (1977). Goal-directed processes in biology. Journal of Philosophy, 74, 261279.Neander, K. (1991). Functions as selected eects: The conceptual analysts defense. Philosophy of Science, 58,

    168184.Nuyen, A. (1989). The Kantian theory of metaphor. Philosophy and Rhetoric, 22, 95109.Philonenko, A. (1977). Lantinomie du jugement teleologique chez Kant. Revue de Metaphysique et de Morale, 82,

    1331.Putnam, H. (2002). The collapse of the fact/value dichotomy and other essays. Cambridge, MA: Harvard

    University Press.Quarfood, M. (2004). Transcendental idealism and the organism: Essays on Kant. Stockholm: Almqvist & Wiksell.Quine, W. V. (1969). Epistemology naturalized. In Quine, Ontological relativity and other essays (pp. 6990). New

    York: Columbia University Press.Quine, W. V. (1975). On empirically equivalent systems of the world. Erkenntnis, 9, 313328.Rang, B. (1998). Zweckmaigkeit, Zweckursachlichkeit und Ganzheitlichkeit in der organischen Natur: Zum

    Problem einer teleologischen Naturauassung in Kants Kritik der Urteilskraft. Philosophisches Jahrbuch, 100,3971.

    Reill, P. H. (2004). Vitalizing nature in the Enlightenment. Berkeley & Los Angeles: University of CaliforniaPress.

    Rescher, N. (Ed.). (1986). Current issues in teleology. Lanham, New York, & London: University Press ofAmerica.

  • J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770 769Richards, R. (2000). Kant and Blumenbach on the Bildungstrieb: A historical misunderstanding. Studies inHistory and Philosophy of Biological and Biomedical Sciences, 31, 1132.

    Richards, R. (2002). The romantic conception of life: Science and philosophy in the age of Goethe. Chicago &London: University of Chicago Press.

    Roe, S. (1981). Matter, life, and generation: Eighteenth-century embryology and the HallerWol debate.Cambridge: Cambridge University Press.

    Roger, J. (1963). Les sciences de la vie dans la pensee francaise du XVIIIe sie`cle. Paris: Colin.Roque, A. (1985). Self-organization: Kants concept of teleology and modern chemistry. Review of Metaphysics,

    39, 107135.Ruse, M. (2000). Teleology: Yesterday, today, and tomorrow. Studies in History and Philosophy of Biological and

    Biomedical Sciences, 31, 213232.Schlosser, G. (1998). Self-re-production and functionality: A systems-theoretical approach to teleological

    explanation. Synthese, 116, 303354.Schurz, G. (2001). What is normal? An evolution-theoretic foundation for normic laws and their relation to

    statistical normality. Philosophy of Science, 68, 476497.Simon, J. (1976). Teleologisches Reektieren und kausales Bestimmen. Zeitschrift fur Philosophische Forschung,

    30, 369388.Sloan, P. (1979). Buon, German biology and the historical interpretation of biological species. British Journal for

    the History of Science, 12, 109153.Sloan, P. (1985). From logical universals to historical individuals: Buons idea of biological species. In J.-L.

    Fischer, & J. Roger (Eds.), Histoires du concept despe`ce dans la science de la vie (pp. 101140). Paris: Singer-Polignac.

    Sloan, P. (1995). The gaze of natural history. In C. Fox, R. Porter, & R. Wokler (Eds.), Inventing human science(pp. 112151). Berkeley: University of California Press.

    Sloan, P. (2002). Preforming the categories: Kant and eighteenth-century generation theory. Journal of theHistory of Philosophy, 40, 229253.

    Sloan, P. (2006). Kant on the history of nature: The ambiguous heritage of the critical philosophy for naturalhistory. Studies in History and Philosophy of Biological and Biomedical Sciences, 37, this issue. doi:10.1016/j.shpsc.2006.09.003.

    Sober, E. (1993). Philosophy of biology. Boulder: Westview.Steigerwald, J. (2003). The dynamics of reason and its elusive object in Kant, Fichte and Schelling. Studies in

    History and Philosophy of Science, 34, 111134.Steigerwald, J. (2006). Kants concept of natural purpose and the reecting power of judgment. Studies in History

    and Philosophy of Science, this issue. doi:10.1016/j.shpsc.2006.09.006.Tucker, A. (2004). Our knowledge of the past: A philosophy of historiography. Cambridge: Cambridge University

    Press.Tuschling, B. (1991). The system of transcendental idealism: Questions raised and left open in the Kritik der

    Urteilskraft. Southern Journal of Philosophy, 30 (Suppl.), 109128.Vartanian, A. (1950). Trembleys polyp, La Mettrie and eighteenth-century French materialism. Journal of the

    History of Ideas, 11, 259286.Wachbroit, R. (1994). Normality as a biological concept. Philosophy of Science, 61, 579591.Wagner, S. (1996). Teleosemantics and the troubles of naturalism. Philosophical Studies, 82, 81110.Wagner, S., & Warner, R. (Eds.). (1993).Naturalism: A critical appraisal. Notre Dame: University of Notre Dame

    Press.Walsh, D. (1996). A taxonomy of functions. Canadian Journal of Philosophy, 26, 493514.Warnke, C. (1992). Naturmechanismus und Naturzweck: Bemerkungen zu Kants Organismus-Begri.

    Deutsche Zeitschrift fur Philosophie, 40, 4252.Wright, L. (1973). Functions. Philosophical Review, 82, 139168.Yovel, Y. (1989). The interests of reason: From metaphysics to moral history. In Y. Yovel (Ed.), Kants practical

    philosophy reconsidered (pp. 135148). Dordrecht: Kluwer.Zammito, J. (1992). The genesis of Kants Critique of judgment. Chicago & London: University of Chicago Press.Zammito, J. (1998). Method versus manner? Kants critique of Herders Ideen in the light of the epoch of

    science, 17901820. Herder Jahrbuch/Herder Yearbook, 126.Zammito, J. (2003). This inscrutable principle of an original organization: Epigenesis and looseness of t in

    Kants philosophy of science. Studies in History and Philosophy of Science, 34, 73109.

  • Zammito, J. (2004a). A nice derangement of epistemes: Post-positivism in the study of science from Quine to Latour.Chicago and London: University of Chicago Press.

    Zammito, J. (2004b). Reconstructing German Idealism and Romanticism: Historicism and presentism. ModernIntellectual History, 1, 427438.

    Zammito, J. (2006). Kants early views on epigenesis: The role of Maupertuis. In J. E. Smith (Ed.), The problem ofanimal generation in modern philosophy (pp. 317354). Cambridge: Cambridge University Press.

    Zanetti, V. (1993). Die Antinomie der teleologischen Urteilskraft. Kant-Studien, 83, 341355.Zumbach, C. (1984). The transcendent science: Kants conception of biological methodology. The Hague: Nijho.

    770 J. Zammito / Stud. Hist. Phil. Biol. & Biomed. Sci. 37 (2006) 748770

    Teleology then and now: The question of Kant " s relevance for contemporary controversies over function in biologyConundrums of lsquo function talk rsquo in biology todayKant and teleologyConclusion: philosophy and science: who is the lsquo Underlaborer rsquo in naturalism?References