Sunteți pe pagina 1din 198

verifiche

Anno XLIII, N. 1-4 Gennaio-Dicembre 2014

Dir. Responsabile: Mario Rigoni • Amministrazione: Casella postale 269 - 38100


Trento - c.c.p. 16677387 • Autorizzazione Tribunale di Trento n. 168 del 13.7.72 •
Spedizione in abbonamento postale gr. IV - Pubblicità inf. al 70% • Composizione e
stampa: Litocenter s.r.l. - Piazzola sul Brenta (PD) - Via G. Rossa, 17 • A. XLIII N. 1-4 2014

The Notion of Organism


Historical and Conceptual Approaches
Edited by Andrea Gambarotto and Luca Illetterati

3 Introduction
Andrea Gambarotto & Luca Illetterati

ARTICLES

15 The Concept of ‘the Organism’ in the Philosophy of Biology


Catherine Wilson

39 Holism, Organicism and the Risk of Biochauvinism


Charles T. Wolfe

59 Organismicity and the Deconstruction of the Organism:


from Substantial Forms to Degrees of Cooperation, Closure and Agency
Georg Toepfer

77 The Rise and Fall of the Machine Metaphor:


Organizational Similarities and Differences between Machines and Living Beings
Victor Marques & Carlos Britos

113 Blumenbach on Teleology and the Laws of Vital Organization


François Duchesneau
137 Teleology beyond Regrets:
On the Role of Schelling’s Organicism in Treviranus’ Biology
Andrea Gambarotto

165 The Concept of Organism in Hegel’s Philosophy of Nature


Luca Illetterati

BOOK REVIEWS

167 S. Normandin, Charles T. Wolfe (eds.), Vitalism and the Scientific


Image in Post-Enlightenment Life Science, 1800-2010, Springer, Dordrecht
2013, 377 pages.
(Andrea Gambarotto)

171 Francesca Michelini, Jonathan Davies (eds.), Frontiere della Bio-


logia: prospettive filosofiche sulle scienze della vita, Mimesis, Milano
2013, 328 pages.
(Elena Tripaldi)

179 Andrea Borghini, Elena Casetta, Filosofia della biologia, Carocci,


Milano 2013, 307 pages.
(Daniele Bertoletti)
INTRODUCTION

by Andrea Gambarotto & Luca Illetterati

In recent years several pieces of scholarship have manifested


a renewed interest in the concept of organism. This interest has
led to an overall reassessment from different perspectives, operat-
ing a fruitful dialogue among historians, philosophers and biolo-
gists1. At the same time, the importance granted to historical and
conceptual approaches has consistently increased. Historians and
philosophers of science have in fact produced important analyses
of the emergence and transformation of the concept, highlighting
the connections between the discussion of organisms originated
in the 18th century and contemporary biological debates2. Within
this framework, the role of Kant appears to be particularly signif-
icant3. Eminent scholars have even portrayed Kant’s view of
organisms as the template of developmental biology and self-

1 P. HUNEMAN, C.T. WOLFE, The Concept of Organism: Historical, Philosophical,


Scientific Perspectives, «History and Philosophy of the Life Sciences», 32 (2-3),
147, 2010.
2 T. CHEUNG, From the Organism of a Body to the Body of an Organism: Occurrence and

Meaning of the Word ‘Organism’ from the Seventeenth to the Nineteenth Centuries, «The
British Journal for the History of Science», 39, 2006, pp. 319–339; P.
HUNEMAN, Métaphysique et biologie. Kant et la constitution du concept d’organisme,
Kimé, Paris 2008, C.T. WOLFE, M. TERADA, The Animal Economy as Object and
Program in Montpellier Vitalism, «Science in Context», Science in Context, 21 (4),
pp. 537-579; C.T. WOLFE, Do Organisms have an Ontological Status?, in P.
HUNEMAN, C.T. WOLFE, The Concept of Organism, cit., pp. 195-132; S.
NORMANDIN, C.T. WOLFE, Vitalism and the Scientific Image in Post-Enlightenment
Life Science, 1800-2010, Springer, Dordrecht 2013.
3 H. GINSBORG, Kant’s Biological Teleology and its Philosophical Significance, in G.

BIRD (ed.), The Blackwell Companion to Kant, Blackwell, Oxford 2006; J. ZAM-
MITO, Teleology then and Now: the Question of Kant’s Relevance for Contemporary
Controversies over Function in Biology, «Studies in History and Philosophy of
Biological and Biomedical Sciences», 37 (4), 748-770; P. HUNEMAN (ed.),
Understanding Purpose: Kant and the Philosophy of Biology, University of Rochester
Press, Rochester 2007.

Verifiche XLIII (1-4), 2014, pp. 3-13.


4 Andrea Gambarotto & Luca Illetterati

organization theory4. Yet, while most recent contributions situate


the notion of organism within the framework of contemporary
philosophy of biology5, the history of concept remains still a
minority option. This special issue aims at partially filling this gap.
In modern science and philosophy «organism» is the word
used to define living beings. However, when we use the term
«organism» with regard to living beings we are not formulating a
tautology, as if the two terms were synonyms. The notion of
organism stresses the fact that living beings are characterized by a
peculiar structure, i.e. organization. More precisely, to use the
notion of organism with regard to living beings means that living
beings are characterized by a peculiar kind of organization in-
volving the relationship between whole and parts. Therefore,
when we say that living beings are organisms, we are claiming that
living beings are characterized by a peculiar relationship between
whole and parts that cannot be found in other entities.
In the transcendental doctrine of method, while he discusses
the conditions for the construction of a system of knowledge,
Kant argues that «without systematic unity, our knowledge cannot
become science; it will be an aggregate, and not a system» 6. Ac-
cording to Kant’s definition, by a system is meant «the unity of
various cognitions under one idea». The system is thus a whole in
which every part receives its own specific place according to a
principle of organization, a whole with a peculiar structure,
according to which all the parts are not merely put side by side,
but have a specific relationship to one another. For this reason,

4 S.A. KAUFFMAN, Investigations, Oxford University Press, Oxford 2000; S.F.


GILBERT, S. SARKAR, Embracing Complexity: Organicism for the 21st Century, «De-
velopmental Dynamics», 129, 2000, pp. 1-9.
5 E.g. T. PRADEU, What is an Organism? An Immonulogical Answer, in P. HUNE-

MAN, C.T. WOLFE, The Concept of Organism, cit., pp. 247-268; M. MOSSIO, A.
MORENO, Organizational Closure in Biological Organisms, in HUNEMAN, C.T.
WOLFE, The Concept of Organism, cit., pp. 269-288; P. HUNEMAN, Assessing the
Prospect of a Return of Organisms in Evolutionary Biology, in P. HUNEMAN, C.T.
WOLFE, The Concept of Organism, cit., pp. 241-372.
6 I. KANT, Kritik der reinen Vernunft (2 Aufl. 1787), in Gesammelte Schrifeten, hrsg.

von der Königlich Preussischen Akademie der Wissenschaften (henceforth


Ak), Bd. III, G. Reimer, Berlin 1911, p. 538.
Introduction 5

Kant claims that in a system «the whole is thus an organism


(articulatio), and not an aggregate (coacervatio); it may grow from
within (per intussusceptionem), but it cannot increase by external
additions (per appositionem). It is, thus, like an animal body, the
growth of which does not add any limb, but, without changing
their proportions, makes each in its sphere stronger and more
active»7. Therefore, the general definition of a system – as a set of
elements characterized by a reciprocal and goal-directed interaction – can be
considered as a possible definition for the notion of organism.
On this conceptual background emerges the idea of the or-
ganism as a key-term to think about different aspects of reality in
opposition to mechanical reductionism. This is the case of the
so-called Lebensphilosophien between the nineteenth and twentieth
century, but also that of systems theory (not only in its biological,
but also in its sociological applications8), politological discussions
(concerning the distinction between individualism and organi-
cism9), and of various organicist approaches in the life sciences10.
These views strive to overcome the classical opposition between
mechanism and vitalism, as they maintain that it is impossible to
explain the organization of living beings by analyzing their ele-
mentary components (molecules, genes). On a molecular level,
organic processes can be fully explained in physical and chemical
terms, but with regard to higher levels of organization it is neces-
sary to acknowledge the existence of emergent features that
cannot be reduced to physicochemical mechanisms11.
The notion of organism refers to a whole characterized by a
peculiar relation to its parts. Therefore, it is defined in opposition

7 Ivi, p. 539.
8 L. VON BERTALANFFY, General System Theory: Foundations, Developments, Applica-
tions, Braziller, New York 1968.
9 A. MEYER, Mechanische und organische Metaphorik politischer Philosophie, «Archiv

für Begriffsgeschichte», XIII, 1969, pp. 128-199; N. BOBBIO, Organicismo e


individualismo, «Mondoperaio», gennaio-febbraio 1983, pp. 99-103.
10 e.g. K. GOLDSTEIN, Der Aufbau des Organismus. Einführung in die Biologie unter

besonderer Berücksichtigung der Erfahrungen am kranken Menschen (1934), neue Auflage


hrgb. von T. Hoffmann und F.W. Stahnish, Wilhelm Fink, Padeborn 2014.
11 Cf. M. BEDAU, P. HUMPHREYS (eds.), Emergence: Contemporary Reading in

Philosophy and Science, MIT Press, Lodon 2007.


6 Andrea Gambarotto & Luca Illetterati

to both aggregates and machines. With regard to aggregates, the


difference is already outlined in classical thought, that is in the
Platonic and Aristotelian distinction between a whole conceived
as to pan, and a whole conceived as to holon. In his Metaphysics12,
Aristotle defines to pan as «those quantities in which the position of
parts does not produce difference», and to holon as «those quantities
in which the position of parts does produce difference». In a
whole conceived as pan the modification of a part does not change
the general structure of the whole, and the absence of a part
produces only a quantitative change, but not a qualitative one. In
the case of a whole conceived as holon the modification of a single
part implies the modification of the whole itself, and is thus to be
seen as a mutilation. Aristotle also outlines the difference between
an organic whole and a machine. Indeed, both artifacts and natural
bodies can form a whole as to holon, because in both cases the
position of parts produces a difference. However, between the
artifact and the natural body there is a difference that cannot be
forgotten: «of this very things – says Aristotle – rather are those
wholes which subsist by nature than such as are made by art»13,
because they are not produced by a simple juxtaposition of parts
put together. Instead, the parts of a living organism are formed
together with the whole, and are not put together by an external
principle. The principle of unity is rather internal to the body itself,
and the parts are nothing more than the articulation of a structure
which is unitary from the beginning.
The relationship between the notion of organism and that
of machine is particularly complex in the early-modern period14.
In early modern thought, in fact, the concepts of organism and
machine often overlap. They both refer to the structure of an
organization unifying a plurality of parts in reciprocal interaction.
It is certainly true that the concept of organism emerges in a
context completely dominated by the idea of nature as a ma-

12 ARISTOTLE, Metaphysics, V, 26, 1023 b 26-1024 a 10.


13 Ivi, 1023 b.
14 Cf. F. DUCHESNEAU, Les modèles du vivant de Descartes à Leibniz, Vrin, Paris 1998.
Introduction 7

chine15. For Leibniz in fact, the term machina does not mean
exclusively a product of the art, but rather a united organization
of distinct parts, in which every part is an instrument (i.e. organon)
in relation to all the others. In this particular case, the notion of
machina is not opposed to that of organism, but rather to aggre-
gates. Of course Leibniz does not ignore the difference between
organized structures in which the origin of organization is exter-
nal to the body and requires the intervention of a designer, and
structures whose organization is spontaneous, but this difference
marks the distinction between artificial and natural entities, not that
between machine and organism. This is clearly testified by the
fact that, in order to indicate structures that are put together
through an external intervention, Leibniz uses the expression
artificial organisms, while to indicate self-organizing structures he
uses the term natural machines. According to Leibniz, the specifici-
ty of organisms is not of being different from machines, but
rather the fact that every part of them is itself a machine, which
is not the case for artifacts. For Leibniz, machine and organism
are not opposed notions, as they both indicate an organized set
of parts in which every part is an instrument (organ) necessary for
the functioning of the whole. The difference is rather between
artificial and natural bodies, because what is artificial always
presupposes an external action, while what is natural acts sponta-
neously, and can thus be called living. But the concept of organ-
ism as such is not different from the concept of machine.
This tension between organism and machine, where these
two concepts are at the same time opposed and overlapping, has
a significant expression in the Kantian thought. According to
Kant, the concept of organism explicitly indicates those products
of nature that cannot be understood by means of an explanatory
framework based on mechanical efficient causality. According to
Kant, to understand living organisms one has to refer to a final
cause, although only in heuristic terms. More specifically, organ-

15 Cf. A.M. NUNZIANTE, Organismo come armonia. La genesi del concetto di organismo
in G. W. Leibniz, Verifiche, Trento 2002; ID., Vita e organismo tra biologia e
medicina: le ragioni di una polemica, in G.W. LEIBNIZ, Obiezioni contro la teoria medica
di Georg Ernst Stahl, Quodlibet, Macerata 2011, pp. 123-194.
8 Andrea Gambarotto & Luca Illetterati

isms imply the reference to the notion of internal purposiveness. This


notion, opposed to that of external purposiveness, is one of the most
original elements in Kant’s discussion of biological issues and it is
what especially makes the conceptual distinction between organism
and machine possible. One speaks of external or relative purposive-
ness when an entity is useful for something else: purposiveness is
thus external because it does not concern things in themselves, but
rather in relation to something else16. One can instead speak of
internal purposiveness when an entity is at the same time «cause
and effect of itself»17, that is when in the end it realizes it is not
external to itself, but rather (the product of) its own realization.
For Kant, organisms exceed mechanical explanations exactly
because they are the cause and effect of themselves in three
different respects: First according to the species, because by pro-
ducing another organism, the latter «continuously preserves itself»
(and is thus at the same time cause and effect of the persistence
of its species); secondly, according to the individual, because in
growth (which «is to be taken in such a way that is entirely distinct
from any other increase in magnitude in accordance with me-
chanical laws») every organism develops itself (and is thus cause)
«by means of material which, as far as its composition is con-
cerned, is its own product» (that is, its effect); thirdly according to
the parts, because the preservation (Erhaltung) of every part «is
reciprocally dependent on the preservation of the other» in a way
in which, as it were, parts are functional to the whole and the
whole is functional to the parts. So that e.g. «the leaves are cer-
tainly products of the tree» (i.e. , its effects), but «yet they pre-
serve it in turn»18 (and are therefore its cause). These characters
are those which, according to Kant, make the distinction between
products of nature and products of art possible. Indeed, alt-
hough a product of art is also an organized body, its purpose is
never internal to itself, but always relative to something else: In
the case of artifacts one always deals with external purposiveness,
and not – as in the case of organisms – with internal purposive-

16 Cf. I. KANT, Kritik der Urtheilskraft, cit., p. 367.


17 Ivi, p. 370.
18 Ivi, p. 371.
Introduction 9

ness. In a machine one part is the instrument for the motion of


another, but not the efficient cause for the production of the
other. Thus, one part of a machine does not produce another
part, and even less does a machine produce another organizing
other matter. In an organism, instead, «each part is conceived as
if it exists only though all the others, thus as if existing for the
sake of the other and on account of the whole»19.
The model of purposiveness employed to account for orga-
nized beings (internal purposiveness) cannot be reduced to the model
of artifacts. If one thinks of the internal purposiveness of living
bodies on the model of the external purposiveness proper to
artifacts, he would exceed the legitimate boundaries of scientific
account. The application of the artifact model to nature would
imply assumptions (such as the idea of an intelligent designer, or
the idea of nature acting intentionally) that a naturalistic account
cannot endorse. Teleological judgment concerning natural bodies,
even if necessary, can therefore only be a reflective judgment and
never a determinant judgment, because otherwise it would imply
the acknowledgment of an intention behind that object. In fact,
according to Kant, a purpose can be explained on the basis of a
designer’s intention, and to think of a purpose without a designer
(as internal purposiveness compels us to do) means thinking of the
purpose without the structure that normally supports it. The
impossibility to grant any constitutive value to purposiveness in
nature – and therefore to consider it only as a maxim of reflective
judgment that has no determining value with regard to living
nature – finds thus its justification in the impossibility, according
to Kant, to think of purposiveness without referring to intention.
In this respect, Kant’s position seems to be marked by a peculiar
tension. On the one hand, Kant operates a distinction between
internal (purposiveness) and external purposiveness, which implies
the idea that the products of organized nature are characterized by a
purpose that makes them irreducible to an explanatory model based
on mechanical efficient causality. By operating this distinction, Kant
seems to think of living beings as irreducible to the way of being of
the artifact. On the other hand, recognizing just a regulative and

19 Ivi, p. 373.
10 Andrea Gambarotto & Luca Illetterati

heuristic value to the principle of internal purposiveness (which is


precisely what characterizes the organism in its difference from
artifacts), he implicitly assumes an artifact model of purposiveness
as a basis for understanding natural purposiveness.
This controversial overlap between organism and machine
plays a role also in recent developments of modern biology. We
speak in fact of artifact model in science every time organisms are
conceived, in terms not far from Kant’s, as if they were intention-
ally organized structures: biology would thus be a science investi-
gating living organisms as if they were artifacts20. This happens
because what makes biology independent from physicochemical
sciences is its peculiar necessity reference to teleological concepts,
which cannot be completely removed21. According to Ernst Mayr
e.g. it is basically impossible to get rid of teleological language in
biology, not just because it is helpful to understand its key-
objects, but because it enables us to express properties that can-
not be expressed without it. Therefore, the basic consequence of
its elimination would be the loss of precious information, that
cannot be recovered using other explanatory models22. In this
respect, biology seems to be incapable of realizing the positivistic
dream and eliminate every teleological language from scientific
explanations. Of course, the idea of teleology as a metaphysical
or theological concept referring to a perfection-directed universe
has clearly been eradicated from modern science. That does not
mean, however, that modern science has completely eliminated
every kind of teleology. On the contrary, according to Ernst Mayr
the use of teleological language is not only legitimate, but also
necessary in order to explain living systems.
On the one hand the notion of organism represents the
word we use to define living beings in opposition to aggregates
and machines. On the other hand, trying to express the purposive

20 Cf. T. LEWENS, Organisms and Artifacts: Design in Nature and Elsewhere, The
MIT Press, Cambridge and London 2004.
21 Cf. W.C. SALMON, Four Decades of Scientific Explanation, Minnesota University

Press, Minneapolis 1989; L. ILLETTERATI & F. MICHELINI, Purposiveness.


Teleology between Nature and Mind, Ontos Verlag, Frankfurt 2008.
22 E. MAYR, The Growth of Biological Thought: Diversity, Evolution, and Inheritance,

Harvard University Press, Cambridge MT 1982.


Introduction 11

organization of parts in relation to a whole, it constantly refers to


an artifact model that seems the only one capable of expressing
the functional relationship of parts with a goal-directed whole.
The first part of the volume includes four contributions
concerned with a broader approach to the notion of organism, its
conceptual history and theoretical implications. Catherine Wilson
addresses the question whether the notion of organism is detec-
tive or projective. In the former case, we can discover whether
particular items are or are not organisms because the category is
in some way mind-independent and theoretically significant for
biology; in the latter case, we decide that certain things are to be
considered organisms because we perceive them as resonating
with our subjective notion of an individual ‘self ’ in a way which
has no theoretical significance for biology. Her conclusion is that
the notion of organism has certain projective features deriving
from human interests and concerns. At the same time, insofar as
some features of what-we-term organisms are theoretically pre-
dictable from the theory of natural selection, she claims that the
concept is grounded in nature. The paper by Charles Wolfe is
concerned with some varieties of biological holism, mainly in the
twentieth and twenty-first centuries. He evaluates some of the
tensions or excessive commitments carried by various forms of
holism, in order to see what can be saved therein. His goal is to
assess the potential legitimacy of a certain kind of holism or
organicism, which, he suggests, needs to steer clear of a variety
of wrongheaded, anti-naturalistic or otherwise excessively falsifi-
able answers to the initial question. The issue is then: which
varieties of organicism are ‘worth wanting’?
Georg Toepfer stresses that organisms provide the best ex-
amples of integrated and functionally closed systems in nature.
However, these systems are not easily specified by one single
criterion or a set of clear-cut criteria. Moreover, all the criteria we
use for the identification of organisms admit of degrees. Organ-
isms seem to be at an extreme point in a whole spectrum of
organized systems in nature. Given this situation, the notion of
‘organismicity’ in the sense of a state variable, i.e. a property of
systems that comes in degrees, seems to be more suitable, at least
in some contexts, for describing the phenomenon than the notion
12 Andrea Gambarotto & Luca Illetterati

of ‘organism’, which denotes individual systems themselves and


not their type of causal structure. Carlos Brito and Victor Marques
focus on the machine metaphor and its meaning for our contem-
porary understanding of living beings. While giving great attention
to contemporary debates, they suggest that the Kantian conceptu-
alization of the organism as a «natural purpose» proves to be the
tool through which several scholars address the question today.
Marques and Brito provide the link to the second part of the
volume, concerned with more specific ʻcase studiesʼ and focused
on the time when biology emerged as a proper field at the turn
of the 19th century. Their general aim is to emphasize the concep-
tual history of the notion of organism as a condition for the
formulation of a properly biological discourse. François Duch-
esneau focuses the attention on the seminal figure of Johann
Friedrich Blumenbach, who is often considered as the principal
initiator of biological vitalism in Germany at the turn of the 19 th
century. This is partly due to the fact that he managed to graft a
hypothesis on the formation of organisms by epigenesis to an
analytic physiology in the style of Haller. This hypothesis attrib-
utes the processes of generation to a formative drive that he terms
‘Bildungstrieb’ or ‘nisus formativus’. This theoretical entity, apart from
meaning a principle of vital self-organization, was further designed
to provide a sufficient reason, a methodological key, for explain-
ing the regular processes that unfold in integrative fashion
through the various functions of organic parts. Blumenbach’s
views played a major role in subsequent biological theorizing.
In his seminal monograph on teleology and mechanics in
eighteenth-century German biology23, Timothy Lenoir claims that
Blumenbach was the first naturalist who accepted the Kantian
understanding of teleological principles and organized it as a
structured research program, later developed by his students at
Göttingen. Lenoir sees his study of the ʻKantianʼ teleomechanis-
tic tradition as an answer to those who wrongly believe that early
nineteenth-century German biology was dominated by Schelling’s
Naturphilosophie. In opposition to Lenoir’s account, Andrea

23T. LENOIR, The Strategy of Life: Teleology and Mechanics in Eighteenth-Century


German Biology, Riedel, Dordrecht 1982.
Introduction 13

Gambarotto argues that Schelling’s organicism played a rather


pivotal role in the formulation of a conceptual framework aimed
at accounting for biological organization. Finally, Luca Illetterati
focuses on the conceptualization Hegel offers of the organism in
his philosophy of nature. The aim of his paper is to show the
naturalistic roots of the notion of subject and to shed light on its
connection to that of organism. Hegel’s position, he argues,
seems to endorse a naturalization of the subject, showing how sub-
jectivity develops primarily in nature and specifically in animals.
The volume concludes with three reviews: Andrea
Gambarotto discusses Vitalism and the Scientific Image in Post-
Enlightenment Life Science (Springer 2013) by Sebastian Normandin
and Charles Wolfe, Elena Tripaldi discusses Frontiere della biologia
(«Boundaries of Biology», Mimesis 2013) by Jonathan Davies and
Francesca Michelini, Daniele Bertoletti discusses Filosofia della
biologia («Philosophy of Biology», Carocci 2013) by Andrea
Borghini and Elena Casetta.
Despite the somewhat different approaches, this collection of
essays can be read as a unitary work. The various contributors
address in fact the same problem, i.e. the role of organismical
notions in biological theorizing, and put the emphasis especially on
the epistemological implications of conceptual history. We thereby
want to suggest that a historical perspective can be a useful tool for
philosophy of biology, one which can help us frame our contem-
porary problems correctly, provide interesting conceptual solutions
and avoid unfortunate cases of theoretical amnesia.
THE CONCEPT OF ʻTHE ORGANISMʼ
IN THE PHILOSOPHY OF BIOLOGY

by Catherine Wilson

Abstract. The notions of ‘selfhood’, self-preservation, and the integration of


the self have a central role to play in moral and political philosophy as well as
in the philosophy of mind. Biologically, however, the notion of the living
individual--the organism-- is problematic. Although we have no difficulty in
picking out, or picking up, or counting individual rosebushes or kittens, it is
difficult to determine where the boundaries of such living things lie and what
belongs to them. The essay discusses the apparent conflict between common
sense and biology, whilst taking account of both Richard Dawkins’s ‘gene-
centered’ perspective and the claims of philosophers who maintain that ‘the
organism’ is an indispensable concept in the life sciences.

Commonsensically, as Michael Ruse points out, there are


cows and horses, oak trees and cabbages. Organisms, living
individuals, special kinds of entities, seem to stand out as figural
against a background realm of rocks, water, air, human artifacts,
dead plants and animals, and the products of organic decay and
decomposition. To ask whether there are ‘really’ organisms might
seem to recall the long-dead debates of the last century over
whether tables and chairs exist as such or are mere illusions
caused by vibrating molecules and reflected photons. This im-
pression is mistaken, for the question persists as an important
one, deepened by the literature of the last decade, both on the
biological side and on the philosophical side.
Ruse’s approach to the question, framed in terms of the is-
sue of reductionism, is a qualified yes: organisms are emergent
entities1. In this paper, I propose to address it from a somewhat
different, though related perspective, by asking whether the
notion is detective or projective. In the former case, we can

1 M. RUSE, Do organisms exist?, «American Zoologist», 29 (3), 1989, 1061-1066,


p. 1066.
Verifiche XLIII (1-4), 2014, pp. 15-37.
16 Catherine Wilson Articles

discover whether particular items are or are not organisms be-


cause the category is in some way mind-independent and theoret-
ically significant for biology; in the latter case, we decide that
certain things are to be considered organisms because we per-
ceive them as resonating with our subjective notion of an indi-
vidual ‘self’ in a way which has no theoretical significance for
biology. My conclusion is that the notion of the organism has
certain projective features deriving from human interests and
concerns. At the same time, insofar as some features of what-we-
term organisms are theoretically predictable from the theory of
natural selection, the concept is grounded in nature. Getting clear
on our contribution and the world’s can effectively resolve the
dispute between philosophers who favor the concept as ontologi-
cally central and those biologists who regard it as arbitrarily im-
posed on the panorama of living nature.
«The concept [of organism]» Richard Dawkins has declared,
is «of dubious utility precisely because it is so difficult to define
satisfactorily»2. Organisms, he maintained in The Selfish Gene, are
not significant units of natural selection; they, as well as the
groups they compose, are like «clouds in the sky or dust-storms
in the desert», «temporary aggregations», «not stable through
evolutionary time»3. Genes, by contrast, said Dawkins, «approach
the idea of indivisible particulateness». His views about the theo-
retical insignificance of the notion are widely shared by some
philosophers of biology. «Even though biology is about organ-
isms, no clear organism concept can be found in biology and
biologists do not seem to have suffered for lack of one», says Jack
Wilson4. The biologist, he goes on to say, is interested in the
evolutionary history of a species and the roles played by «chance,
homology, and selection, and design [constraints]» in that history.
For most biologists, biology is about organisms in the same

2 R. DAWKINS, The Extended Phenotype, Oxford University Press, Oxford and


New York 1982.
3 Ivi, 34-5.
4 J.A. WILSON, Ontological Butchery: Organism Concepts and Biological Generalizations,

«Philosophy of Science», 67, 2000, 301-311, p. 301. See also J.A. WILSON,
Biological Individuality: The Identity and Persistence of Living Entities, Cambridge
University Press, New York 1999.
Articles The Concept of Organism in the Philosophy of Biology 17

sense that physics is about material objects. While there are theo-
retically deep things to be learned about (the things we call)
organisms, there is no theoretically deep account of what it is to
be an individual organism. Physics should be able to explain why
some material object is solid, brown, and weighs 20 kg. Pursuing
such explanations led physicists into the heart of the theoretical
disciplines of optics, mechanics, and the theory of forces, but the
term ‘material object’ has only a vernacular, domain-indicating
and not a physics-theoretical sense. It is a term that entered the
philosophical vocabulary at a historical moment to generalize
over planets, rocks and rolling balls, and to contrast such items
with animate, self-moving beings. Similarly, biology should ex-
plain how organisms grow, succumb to or resist disease, repair
damages, interact, reproduce and die, and these inquiries led
biologists into physiology, molecular biology, and ethology and
other fields of the life sciences. But the term ‘organism’ might
have only a vernacular, domain-indicating sense – a sense to
which I will incidentally adhere in the remainder of this essay.
Biologists did not discover that there were organisms in the same
way those physicists discovered that there were electrons, or even
atoms. Rather, the term came into use at about the same time as
the term ‘material object’ came into use in philosophy in order to
be able to generalize over plants, animals, and protozoa and to
contrast such entities with artificial machines and with natural
things that exhibited growth such as crystals and with other
mineral items. The term «organism» has a philosophical history,
one shaped by theological, socio-political and even existential
concerns, as well as by observation and experiment in zoology
and botany, medicine, physiology, and related fields. It seems to
reflect our concerns with the way in which we interact with the
rest of the world, and the way in which organisms interact with
one another in ways of interest or relevance to us. It is (remotely)
conceivable that technological developments – the creation of
artificial life and the proliferation of bionic people and animals
will erase the term from our commonsense vocabulary in centu-
ries to come.
To defenders of the theoretical importance of the concept,
neither the historicity of the term nor its psychological resonance
18 Catherine Wilson Articles

impugns its explanatory role. Evan Thompson maintains that «to


account for certain observable phenomena, we need the concepts
of organism (in the Kantian sense of a self-organizing and imma-
nently purposive whole)»5. According to Dennis Walsh, protest-
ing against the anti-individualism’ of contemporary biology, the
very notion of an evolutionary history requires essential refer-
ences to organisms:

Genes or traits…don’t occur as disembodied members of popu-


lations; they occur as constituents of organisms. Likewise, the
fate of a gene or trait is tied to the fate of the organism it is in.
Consequently […] gene fitnesses […] are realized by the differ-
ential survival and reproduction of organisms in which gene
(trait) tokens occur6.

This is an intriguing claim, and it is the purpose of the paper to


try to explicate the conflict between organism-skeptics and organ-
ism-defenders like Thompson and Walsh. My conclusion, as
foreshadowed above, is that both sides are partially right.

1. The Biologists’ Perspective

Most of life, one skeptic remarks, does not fit preconceived


notions of countable individuality. «It is apparent», Leo Buss
comments, ‘that individuality is a derived character, approximated
closely only in certain taxa. This fact is of substantial interest for
it means not only is it inaccurate to consider the individual as the
sole unit of inheritance in most taxa, but also that we have little
assurance that it is appropriate to assume this to have been the
case throughout geological time»7. Many living things – Jack
Wilson cites «colonial siphonophores, slime molds, sponges,
gigantic fungi and lichens» are neither obviously individual organ-

5 E. THOMPSON, Mind in Life: Biology, Phenomenology and the Sciences of Mind,


Harvard University Press, Cambridge MA 2007, p 164.
6 D. WALSH, Evolutionary Essentialism, «British Journal for the Philosophy of

Science», 57, 2006, 425-448, p. 435.


7 L. W. BUSS, The Evolution of Individuality, Princeton University Press, Princeton

1987, p. 20.
Articles The Concept of Organism in the Philosophy of Biology 19

isms nor obviously not8. A slime mold forms a body called a grex
from free-living single cells. The grex resembles a slug, respond-
ing to light and moving in co-ordinated fashion; it raises itself
into a tower, emits spores and disperses9. A bee colony in which
only one female is capable of reproduction, the others function-
ing either to provide sperm or to provision the colony and its
offspring, can be viewed either as a collection of organisms or as
a single unit.
Peter Godfrey-Smith reviews a panorama of life forms that
do not fit the stereotype of the individual organism. These in-
clude aspens, which spread via an underground root system to
produce a set of physically connected trees, and other plants like
strawberries and violets that spread by runners and may or may
not remain connected to the parent plant, and dandelions, which,
though we dig them up one by one, are members of a single
clone. The picture of the organism as individual is further eroded
by symbiosis; optically separate individuals with their own repro-
ductive lineages may live entwined, or side by side, or one within
the other in a state of interdependency. These include the mito-
chondria, once free living bacteria, and such curious partnerships
as the shrimp and fish which live in symbiotic pairs and ‘whose
heads may be seen poking out of the same burrow10. Metamor-
phosis, in which the cells of the worm-like larva die and are
consumed for nutrition as the imaginal disks of the fly begin to
develop, violate the commonsense conception of an organism as
having a basic shape, a basic mode of life, and only one death.
The caterpillar-butterfly effectively dies twice.
Further, for the biologist, there is no principled way of de-
ciding what belongs to the organism and what it has produced or
made. The mollusk’s shell is firmly attached to the rest of its
body, whilst the fox’s burrow is not. The oyster does not need
eyes or a brain to make its protective covering while the fox does.
8 J. WILSON, Ontological Butchery, cit., p. 302.
9 Ivi, p. 304.
10 P. GODFREY-SMITH, Populations and Natural Selection, Oxford University

Press, Oxford & New York 2009, pp. 70-75. A number of these examples
were originally cited by D.H. JANZEN, What are Dandelions and Aphids?, «Ameri-
can Naturalist», 111, 1977, pp. 586-9.
20 Catherine Wilson Articles

But both structures have determinate forms and similar protec-


tive functions. Does it matter that the fox used parts of its body,
namely eyes and brain, to produce its burrow rather than just
secreting it? Turner argues that animal-built structures such as
burrows, galls, hives, mounds, and caddis-fly houses are physio-
logical structures, whose physical properties, shaped by natural
selection, can be analyzed in great detail11. The colorfully-
decorated bower of the bowerbird, as John Bonner points out, is
not really a nest, but a form of feathering, a display feature that is
a little safer for the bird than feathers growing out of its body
would be, and that it will construct whenever materials are availa-
ble12. As advocates of the importance of the concept of ‘niche
construction’ point out, an organism poorly adapted to a new
environment can respond by changing its habits as well as chang-
ing its physiology; both are accordingly features of its phenotype.
Thus earthworm kidneys are said by Odling-Smee et al. to be
poorly adapted to a terrestrial environment: their tunneling be-
haviour compensates, enabling them to regulate their fluid bal-
ance13. We can take the bowerbird out of its bower, or the earth-
worm out of its tunnel, or even the oyster out of its shell. The
one cannot then reproduce by its usual means, the others cannot
live by their usual means, though they could with extra help or
even – given an ingenious variant and the right materials – by
forming a new habit, as their ancestors must have done.
Further still, the organism’s morphology and physiology and
its behaviour may reflect the preferences and powers of other
organisms and may have evolved to serve their reproductive
ends, not its own. The song of the male canary, as Dawkins
observes, can unleash the cascade of hormonal events necessary
to induce the female canary to mate, whether the time is ideal for
her or not14 The warblers that raise the cuckoo’s young have

11 J.S. TURNER, The Extended Organism: The Physiology of Animal-Built Structures,


Harvard University Press, Cambridge MT 2000.
12 J.T. BONNER, Life Cycles: Reflections of an Evolutionary Biologist, Princeton

University Press, Princeton 1993, p. 188.


13 F. ODLING-SMEE, K. LALAND, & M.W. FELDMAN, Niche Construction: The

Neglected Process in Evolution, Princeton University Press, Princeton 2003, pp. 11-12.
14 R. DAWKINS, The Extended Phenotype, cit., p. 63.
Articles The Concept of Organism in the Philosophy of Biology 21

evolved to raise whatever young are in the nest, but not specifi-
cally to raise young cuckoos. Whenever morphology appears
typical or behavior appears stereotypical, we should assume that it
is caused by replicator forces, but not necessarily forces acting
from within the organism exhibiting the form or the behavior.
Sneezing may not be a defensive reaction of a rhinovirus’ host,
but rather a means for spreading the virus contrived by it.
Within its ecosystem, each plant or animal is a necessity, a
condition of life, for certain other life-forms, and a hostile pres-
ence, undermining the conditions of life for still others. It is a
loose federation of other plants and animals, both those neces-
sary to it and those dependent on it. We are pervaded by other
living beings; without our gut bacteria we could not digest our
food; without sebum-devouring scavengers on our skins, we
would break out in pustules. Metchnikoff discovered that the
phagocytes that primitive organisms use for nutrition were turned
to a new use in organisms with digestive cavities, namely that of
consuming senescent, malignant, defective, damaged, and dis-
eased cells’ or pathogens15. They were likely once free-living,
scavenging cells that teamed up with more complex structures to
work for them. In pregnancy, fetal cells circulate in the mother’s
body rendering her ‘chimerical’ and perhaps providing a long
lasting resource of genetic material for repair of the body in
females, as well as having other less favorable effects 16. To quote
from a recent abstract:

An increasing number of studies have recently detected within-


organism genetic heterogeneity suggesting that genetically ho-
mogeneous organisms may be rare [and have implications for]
the fitness of the individual. The costs of IGH (intraorganismal
genetic heterogeneity) include cancerous growth, parasitism,
competitive interactions and developmental instability, all of
which threaten the integrity of the individual, while the potential
benefits are increased genetic variability, size-specific processes,
and synergistic interactions between genetic variants. The par-

15 A. TAUBER, Metchnikoff and the Phagocytosis Theory, «Nature Reviews: Molecular


Cell Biology» 4, 2001, pp. 897-901.
16 D. LISSAUER et al., Persistence of Fetal Cells in the Mother: Friend or Foe?, «British

Journal of Obstetrics and Gynecology», 114, 2007, pp. 1321-25.


22 Catherine Wilson Articles

ticular cost or benefit of IGH in a specific case depends on the


organism type and the origin of the IGH 17.

Immunology has been looked to in the hope of deriving


from the phenomena a robust conception of selfhood: for even
primitive animals such as sponges and jellyfish mount defenses
against pathogens and intrusions such as splinters; the animal
appears to discriminate between self and other18. Yet here too we
find latent disunity accompanying phenomenological unity:
sponges attempt to wall off diseased tissue and to combat aggres-
sive bacteria, but tolerate symbiotic bacteria forming up to 40%
of their masses19 Co-operation, surveillance and self-maintenance
are transitory and imperfect processes. On one hand, there are
biologically tolerable materials that do not generate an inflamma-
tion reaction but are not part of the organism such as titanium
implants; on the other hand, malignant cells of its own can kill
the organism. In autoimmune disease, elements that ought to be
recognized as the organism’s own become the object of attack.
To appreciate the difficulty in giving a scientifically accurate
characterization of ‘the organism,’ consider Dawkins’s own at-
tempt to respond to charges that he has ‘lost’ it. For Dawkins, the
explanandum for the biologist is an array of ‘phenotypic characters;’
the explanans consists of replicators and their activities, and organ-
isms drop out of the picture as theoretically significant entities:

I have a dim vision of phenotypic characters in an evolutionary


space being tugged in different directions by replicators under
selection. It is the essence of my approach that the replicators
tugging on any given phenotypic feature will include some from
outside the body as well as those inside it. Some will obviously
be tugging harder than others, so the arrows of force will have
varying magnitude, as well as direction. My hunch is that almost

17 M. PINEDA-KRCH & K LEHTILÄ, Costs and Benefits of Genetic Heterogeneity within


Organisms, «Journal of Evolutionary Biology», 17, 2004, pp. 1167-1177, p. 1167.
18 See the discussion and defense of the immunological criterion in Ellen

Clarke, The Problem of Biological Individuality, «Biological Theory», 5, 2010, pp.


312-325.
19 R. OSINGA et al, Sponge–Microbe Associations and their Importance for Sponge

Bioprocess Engineering, «Hydrobiologia», 461, 2001, pp. 55-62.


Articles The Concept of Organism in the Philosophy of Biology 23

all phenotypic characters will turn out to bear the marks of con-
flict between internal and external replicator forces 20.

Nevertheless, Dawkins makes an attempt to characterize the


vernacular notion, focusing on physiological uniformity and
functional unity as J.S. Huxley had much earlier. He cites the
following features of organisms21

(1) An individual organism is walled off from the environment


(2) An individual organism has a complete set of the same genes
in every cell in its body
(3) It is immunologically self-recognizing, self-accepting and
other-rejecting
(4) It is a behavioral unit, with a CNS and co-ordinate limbs that
‘takes decisions’ as a unit.

But the entities we call organisms – you, me, the sheep in the
field, the rosebush – do not, as the foregoing has made clear, fit
these criteria in the strict sense of necessary and sufficient condi-
tions required by philosophers, as Dawkins himself is well aware.
We are mostly walled-off, mostly self-recognising, mostly able to
take decisions as a unit.
J. Wilson argues that biological individuality decomposes in-
to a multitude of restricted notions, including spatio-temporal
continuity, developmental history, and functional unity22. These
are conspicuously united in the higher animals, the paradigmatic
organisms, but the class so identified depends on the grid we
have imposed. Godfrey-Smith suggests in turn that we can
achieve a form of reflective equilibrium by adjusting our com-
monsense notion of the individual organism to values on three
distinct parameters. These parameters include the Dawkinsian
notion of a bottleneck – is there an interval in the life-cycle of the

20 R. DAWKINS, The Extended Phenotype, cit., p. 248.


21 Ivi, p. 254; see J.S. HUXLEY, The Individual in the Animal Kingdom (1912) repr.
Oxbow Press, Woodbridge CT 1994. See also the related but amplified list of
features in J. WILSON, Biological Individuality, Cambridge University Press,
Cambridge 1999, p. 56.
22 Ivi, p. 60.
24 Catherine Wilson Articles

entity when it consists of a single cell capable of mutation? integra-


tion – can its parts live separately or do they depend on one an-
other? and separation – is there separation of somatic cell lines
from the germ line? The three dimensional space that these
parameters determine is occupied in every region by living things.
Paradigmatic organisms thus have bottlenecks, are well integrat-
ed, and separate their gametes, which in humans contain half the
usual number of chromosomes, from the cells with a full com-
plement23. Each of the parameters is theoretically rich in its own
right. But it cannot be said that this ingenious proposal captures
the intension of the term ‘organism’ as it is used by speakers of
English, even if it captures its (vague) extension. It is as though
someone were to give a set of physico-chemical specifications,
with Boolean complexity – many ‘ands’ and ‘or’, that applied to
all gemstones and forced us to include a few stones not normally
to be found in jewelers’ windows, but that fulfilled the criteria
anyway. The definition would be extensionally correct, but not
reveal the meaning of the term ‘gemstone,’ which requires expli-
cation in terms of human values and practices.
This is not an objection to the procedure, for Godfrey-Smith
appears to side with the biologists who do not find the term
theoretically useful; with so many entities functioning as the
participants in Darwinian selection, there is no more need for
Godfrey-Smith than for Dawkins to award special status to ‘the
organism’, though for entirely different reasons. The notion of a
‘gene’, he thinks, is too vague and imprecise to function as a
fundamental theoretical concept. A gene is simply a ‘region’ on
the chromosome susceptible of crossing over to the homologous
chromosome, not a definite, countable thing24. Dawkins’s insist-
ence that genes have the kind of identity and permanence that
makes them the unique units of selection is, he thinks, mistaken; a
more appropriate candidate for a fundamental notion is that of a
‘Darwinian population’ as comprising an array of units anywhere
from DNA fragments to cell lineages to large breeding groups.

23 P. GODFREY-SMITH, Populations and Natural Selection, cit., p. 95.


24 Ivi, p. 136.
Articles The Concept of Organism in the Philosophy of Biology 25

2. The Philosophical Concept of the Organisms

Despite the considerations just cited, it appears that we can-


not help conceptualizing organisms as self-sustaining living indi-
viduals, thoroughly unlike mere physical objects. Thompson
appeals to the notion of autopoiesis or self-construction and self-
maintenance, which Francisco Varela originally described as the
essential property of the living cell. The organism, he says, has a
boundary, a semi-permeable membrane. It stands out of the
molecular soup of the environment as a distinct figure. There is
interaction at the boundary with the environment but the organ-
ism is selective about what can get out and what is allowed in.
Within this boundary, the organism produces and maintains itself
‘autopoetically’. It creates and repairs the components necessary
for it to carry on with creating and repairing itself. Thompson
refers in this connection to an «inescapable, pragmatic circulation
and mutually constraining relation between science and experi-
ence»25. Life, he declares, can only be known by life; concepts
such as organism «are available only to a bodily subject with
firsthand experience of its own bodily life». The organism, he
maintains, appears from a «transcendental perspective»; it is not a
ready made category out there in the world. Our subjective expe-
rience and only our subjective experience enables us to extract an
adequate characterization of the distinctive and defining features
of this ontological kind. The source for the meaning of the con-
cept of the organism, he says,

is the lived body , our original experience of our own bodily ex-
istence…These concepts are not derivable from some observer-
independent, nonindexical, objective physico-chemical descrip-
tion […]. [They] are available only to a bodily subject with
firsthand experience of its own bodily life26.

The concept is accordingly both detective and projective in


the eyes of its major proponents. It is detective, insofar as it is a

25 E. THOMPSON, Mind in Life, cit., p. 164; cf. F.J. VARELA, Principles of Biological
Autonomy, Elsevier, New York 1979.
26 Ibidem.
26 Catherine Wilson Articles

feature of reality that it contains some boundaried, autopoietic,


cognizing, reproducing entities. It is projective, insofar as only
our embodied experience enables us to grasp what an objective
disembodied consciousness, attuned to the presence of physico-
chemical events occurring in a determinate spatio-temporal
framework could not. We are aware, one might reflect along
these lines, that we have a boundary –our skin-- that it is painful
and dangerous to violate, and we aim to maintain it in good
condition. The spontaneous growth of hair and fingernails and
the healing of wounds inform us that the processes of renewal
and repair come from within us. We are accustomed to getting
sensory information from the environment that guides actions of
approach and avoidance. We know what it is like to send into the
world things that are our own productions and that bear the
marks of the producer upon them.
It is not clear however that the term ‘organism’ is more tran-
scendental than the terms ‘material object’ or even ‘stuff,’ terms
indicating items in our experience that lack the features of auto-
poiesis or boundaries or both. It is not within the scope of this
paper to give a history of the term and its applications, but it is
worth mentioning how questions of agency, mechanism, teleolo-
gy, generation, and mortality, shaped the concept27. Aristotle’s
notion of the living or ensouled thing did not distinguish between
plants and animals and other beings such as the stars, and miner-
als and crystals were not definitively assigned to the inorganic
realm until the end of the 18th century. The notion of the living
individual became the focus of attention in the debates over the
coherence of the corpuscularian philosophy and the immortality
of the soul of the late 17th century. Leibniz, for example, insisted
on an ontological distinction between organisms, which he con-
sidered to be immortal, and perishable aggregates, and indeed
made that distinction the foundation of his metaphysics28.

27 See also J. WILSON, Biological Individuality. The Identity and Persistence of Living
Entities, Cambridge University Press, Cambridge 2007, pp. 28-32.
28 G.W. LEIBNIZ, Letter to Arnold, October 9, 1687 tr. in Philosophical Papers and

Letters, 2nd ed., ed. L.E. Loamier, Reidel, Dordrecht 1969, pp. 338-348; pp.
342-5.
Articles The Concept of Organism in the Philosophy of Biology 27

Hobbes’s premise that self-preservation or ‘conservation’ is the


principal concern of the human being is the foundation of his
political philosophy29, and, while Spinoza deems the individual
only a mode of a single, universal substance, the fortification of
the self against the assaults of the environment that disturb tran-
quility is the major theme of his ethics. Finally, Kant’s denial that
there would ever be a Newton of a blade of grass 30 expressed his
aversion to the notion that the mechanical philosophy might
eventually provide a complete account of life that would under-
mine the significance of personal agency and responsibility. Un-
like their ancient predecessors, the early moderns did not grasp or
at least did not emphasize the significance of reproduction as a
defining feature of the living thing. Even the writings of Darwin
and the early Darwinists give the survival of the favoured indi-
vidual – the faster, stronger, cleverer one – a theoretical promi-
nence it does not possess in contemporary biology, in which any
trait may or may not be conducive to individual survival and
individual survival is only important insofar as it makes reproduc-
tion possible or enhances reproductive success.
The metaphor of human life as a struggle by individuals to
maintain their integrity, to compete for resources, and to preserve
themselves against external destructive forces is a theme of meta-
physics and political philosophy that is historically prominent, but
that antedated much real understanding of life. It is salient for us
that organisms co-operate and compete, that they try to protect
themselves against marauding others and succeed for a time, but
are eventually overcome by them – death from cancer, or infec-
tion, or ingestion, or crushing or wounding being the usual route
to nonexistence. It is also salient for us that we have an integu-
ment which is painfully and dangerously breached, and an ability,
though an imperfect one, to repair damaged components. The
macroscopic category of ‘animal’ as a kind of intentional agent

29 T. HOBBES, Leviathan (1651), ed. Richard Tuck, Cambridge University Press,


Cambridge 1996, pp. 87-8.
30 I. KANT, Critique of Judgement (1790), tr. Werner S. Pluhar, Hackett, Indianap-

olis 1987, p. 400. (Cited after volume V of Kant, Gesammelte Schriften, ed.,
Akademie der Wissenschaften, Reimer, Berlin 1912).
28 Catherine Wilson Articles

has significance for us, and this significance is wired in by evolu-


tion. Studies of change blindness indicate that animals are pro-
cessed differently by the human visual system than mobile arti-
facts, even dangerous ones such as cars31. They are noticed. This is
said to reflect a special human sensitivity to agents, entities with
purposes that are friendly and hostile, and that bear attentive
watching to see what they want and what they are going to do
next. Plants, as well, stand out from the environment as growing
and reproducing entities that, like animals, begin very small, grow,
succumb to diseases and die. To the extent that we can perceive
an amoeba or a mushroom as similar to the paradigm, we are
satisfied to call it an organism, but we draw the line at a spot of
mildew. Where projection fails, as in the case of certain slimes,
fungi, colonies, or symbiotic partners, we fail to ‘see’ a single
living individual, as we also fail to do when confronted with
immature parts behaving somewhat like mature wholes. We have
great difficulty in deciding whether a chicken egg in which the
tiny punctum saliens has become visible is an organism and whether
a tree branch that has been put into a glass of water and is now
sprouting roots or opening its buds is an organism or simply a
metabolizing entity. But we can agree that nothing whatsoever,
from the scientific perspective, hangs on the decision.
The early modern construction of the notion of the organism
explains the historically-conditioned focus on self-preservation,
autonomy, and separateness as essential characteristics of the
organism, as well as the relative indifference to reproduction in
these classical formulations. When organisms are characterized as
bounded, persisting, struggling entities, whose capabilities require
explanation in holistic terms, or elude understanding in the lan-
guage of molecular interactions, one senses the pull of metaphys-
ics and ethics and their particular concerns, concerns to which
practicing biologists are typically indifferent. Walsh’s challenge,
by contrast, recognizes the centrality of reproduction in the
theory of life, and this lends it a special interest.

31J. NEW, L.COSMIDES, & J. TOOBY, Category-Specific Attention for Animals Reflects
Ancestral Priorities, not Expertise, «Proceedings of the National Academy of
Sciences», 104, 2007, p. 40.
Articles The Concept of Organism in the Philosophy of Biology 29

3. Organism as a Theoretically Required Concept

Walsh employs several arguments for the theoretical rele-


vance of the concept in question. He maintains that reference to
a phenotypic entity that is both mutable and stable is necessary
for evolutionary change to occur, and that organisms are those
entities in nature that enjoy a balance between mutability and
stability, thanks to their possession of plasticity and corrective
mechanisms32. As this argument is rather indirect, I will concen-
trate on his more direct argument for the indispensability of the
concept33 which I reconstruct more formally below.
Role Premise: Biologists may be called on to explain why
populations have historically evolved in the directions they did,
i.e. why gene frequencies in those populations changed over
time, resulting in changes to phenotypical characters. They
should be able to tell us why horses grew larger after the Eo-
cene, why animals isolated on islands tend to dwarfism, and
why species that migrated to the Arctic developed pale fur.
Ontological Premise: Long life, health, and reproductive suc-
cess can only be enjoyed by individual organisms, not by mere
replicators such as genes or populations. Conclusion: for biol-
ogists to fulfill one category of explanatory tasks that can be
legitimately assigned to them, they must make essential refer-
ence to organisms as living individuals.
In response, the organism-skeptic who maintains that the
notion is purely projective might deny the Role premise. She
might insist that explaining why a breeding population devel-
oped new and shed old characteristics is not the professional
work of practicing biologists. To the extent that biologists do so,
they are indulging in a fun and popular recreational activity, that of
thinking up unverifiable just-so stories. From the gene’s eye per-
spective, the mosaic of genes active over the entire earth changes
its composition from instant to instant, but not for particular
‘reasons,’ any more than, from the perspective of quantum me-
chanics, fruit ripens and falls to the ground for particular reasons.

32 D. WALSH, Evolutionary Essentialism, cit., p. 436.


33 Ivi, p. 435.
30 Catherine Wilson Articles

The defender of the ‘detection’ view will insist that the fields
of ecology and animal ethology, if not laboratory biology and the
mathematical branches of the study of selection, must recognise
the phenotype that forages and mates as a unit, when they are not
dealing with the problematic cases of certain plants, slime molds,
and siphonophores, and that the colonial or chimeric qualities of
an organism are no barrier to doing so. Nor is the fact that selec-
tion may occur at other levels than the organismic. Further, not
all proferred explanations of evolutionary change need to be
regarded as just-so stories; experimental intervention that alters
the qualities of the phenotype can tend to the confirmation of
hypotheses, as when, for example, birds are outfitted with hats or
other adornments to test their attractiveness to the opposite sex34.
Other interventions could well indicate the significance of traits
for survival to maturity or reproductive success, subject always to
the usual problems faced by statisticians of sample size and un-
known co-factors.
Further, Dawkins’s reference to organisms as ‘vehicles’ for
genes is somewhat misleading. This is obvious from the fact that
if a box of genes were driven around in a car, there would be no
evolutionary change as we understand the term, even if they
engaged in replication. Vehicles are not merely repositories, but
genuine enablers, through their qualities and their behaviour, in a
world of other enablers and disablers. Their vehicles must furnish
the energy that enables genes to carry out their physico-chemical
labors, as well as providing them with the morphological struc-
tures that enable them to get into the next generation. Bodies,
birth, nutrition, and some form of reproductive anatomy and
physiology are thus essential conditions of evolutionary change: a
given phenotype is a particular way of realizing these conditions.
Real-world conditions then put flesh on the bones of the abstract
vehicle, endowing vehicles with boundaries, life-cycles, appetites,
growth from a smaller form or bottleneck, and with homeostatic
dispositions, and autopoietic abilities, elaborating upon and diver-
sifying phenotypes. These elaborations are salient to us, for repre-

34 KLAUDIA WITTE & EBERHARD CURIO, Sexes of a monomorphic species differ in


preference for mates with a novel trait, «Behavioral Ecology», 10 (1), 1999, pp. 15-21.
Articles The Concept of Organism in the Philosophy of Biology 31

sentatives of our own phenotype not only observe and experience


these features of themselves, they seeks to enhance them by
artificial means, from thermostats and fever-reducers for help
with homeostasis, wound-plasters for maintaining the boundary,
parasite-expellers and acne remedies for destroying invaders,
artificial limbs for those we cannot regrow, and incitements to
compete for and accumulate resources to facilitate nourishment
and reproduction.
Our organism-skeptic, having accepted that explanation of
evolutionary change, to the extent that it is possible, is someone’s
responsibility, may argue that our projective tendencies will blind
us to many significant drivers of change. Focusing on what look
to us like organisms, and on what we human beings consider
their salient traits—size, beauty, swiftness, disease resistance—
because these seem to be important in our selective activities,
especially those involving ‘mate choice’, blinds us to the contribu-
tion of competition between genes, parasitism at the genetic level,
and a host of other subtle factors that bring about alterations in
the vast mosaic of the terrestrial gene pool35.
The very heterogeneity of ‘the organism,’ however, becomes
scientifically intelligible to us through our employment of vernac-
ular concepts borrowed from human social life. Without projec-
tion from our experience at the ‘vehicular’ level, there can be no
understanding of the insights achieved by evolutionary theorists.
This point is developed in the next section.

4. Phenomenology and Value

The consequences of our philosophical heritage and our psy-


chological endowments are that we – the disunified, extended,
genetically heterogeneous organism of the genocentric view – see
ourselves nevertheless as individual organisms. A disembodied
intelligence that did not have life experiences like ours would, for
this reason, have a great deal of difficulty understanding the basic

35 For accusations of explanatory folly, see J. FODOR & M. PIATTELLI-


PALMARINI, What Darwin Got Wrong, Farrar Strauss and Giroux, New York
2010.
32 Catherine Wilson Articles

concepts of Dawkins’s Extended Phenotype. For the organism is


the very thing that is extended, parasitized, assisted, or ham-
pered in its functions. It is the thing that will fail to complete its
life cycle if deprived of certain of its parts and products. What is
counter-intuitive on one level – the disunity of the organism and
its indefinite spatial boundaries – is fully intuitive on another.
Dawkins’s presentation of his world of replicator forces is
intellectually compelling because we have first hand embodied
experience, not just of selfishness, but of parasites, harmless
hangers-on, internal conflict in friendly teams, and manipulation
by and of others. We know what it is like to consider cars,
houses, literary productions or clothing extensions of ourselves.
Psychologically, we know what it is like to be a field in which
multiple conflicting forces are at play.
Organisms, then, are those things in whose life or death we
take an interest, and which we regard as being like us in certain
phenomenological respects – having shapes, skins, being capable
of repair, taking in nutrition and so on. We, with our evolved and
useful sense of a self with goals, impediments, helps, and setbacks
confront the varieties of life and select organisms from colonies,
cells, organs, and blobs of protoplasm in a way that reflects our
characteristic forms of engagement with our neighboring life
forms. This is the projective aspect of the concept. Yet it is uni-
fied by two features; our identification of and with other organ-
isms reflects a genuine appreciation of their distinctive features.
For the possession of boundaries and autopoiesis are traits – ueber-
traits – shared by many living forms on account of their common
descent, if not from a single ancestor, at least from a very small
number of viable forms of ‘proto-life,’ within the constraints
imposed by the earth’s atmosphere, its gravity, illumination, and
the history of disasters and upheavals it has passed through that
led to massive extinctions. They are traits that facilitate survival
and reproduction under a wide variety of circumstances. Further,
stable assemblies can be formed more easily and so are more
likely to form spontaneously from pre-existing, stable sub-
Articles The Concept of Organism in the Philosophy of Biology 33

assemblies than de novo36. So organisms are federations of organ-


ism-like entities which may themselves be only loosely integrated,
and they in turn form parts of larger organism-like entities.
The early modern conception of the organism as a well-
circumscribed, self-maintaining and self-repairing locus of power
and vulnerability was accordingly a discovery of some important
features of plants, animals, and microorganisms. At the same
time, it was projective, a transcription of an emerging introspec-
tive or reflective sense of the self as a fortified individual, deserv-
ing of recognition and respect, and entitled to further protection.
The integrated and well-defended self underwrote and continues
to underwrite a set of ethical stances and social movements, from
rugged individualism and self-ownership37 to human rights, to
anti-vaccination campaigns38 to anti-feminism39, to vegetarianism.
Moral integrity and entitlement are seen as closely related to
physical inviolability.
The biologists are accordingly right to imply that we human
beings possess an exaggerated view of the unity and integrity of
the self that is derived by way of idealization from the practices of
the care of the self typical of our phenotype. The need for these
practices testifies to the ways in which we fail at being perfectly
homeostatic, bounded, and autopoetic, and the ways in which our
phenotypes are interwoven with phenotypes produced by other
collections of genes. The notion of the organism thus reflects to
some degree an illusion we hold about ourselves and about the
rest of nature to the extent that we regard it as a collection of

36 H. SIMON, The Architecture of Complexity, «Proceedings of the American


Philosophical Society», 106, 1961, pp. 467-482.
37 Robert Nozick is recognized as defending what he calls in Anarchy, State and

Utopia (Basic Books, New York 1977) «the classical liberals notion of self-
ownership» p. 172 et passim against attempts to appropriate or divert the self ’s
productive powers.
38 Kant, proclaiming that we do not own ourselves, insists that we have a duty

to maintain bodily integrity that prohibits organ donation. See Lectures on


Ethics, tr. by P. Heath, Cambridge University Press, Cambridge 1997, p. 371.
39 Women may be perceived as leakier and more permeable, as possessing less

physical integrity than men. See M. SCHILDRICK, Leaky Bodies and Boundaries:
Feminism, Postmodernism and (Bio)ethics, Routledge, London 1997.
34 Catherine Wilson Articles

discrete and self-sufficient individuals, ignoring the plenitude of


life forms that do not resemble us, as well as the finer details of
our own constitutions and surroundings. At the same time, the
philosophers are right to point out that vehicles have taken on
the characteristics ascribed to organisms by philosophers for
good reason, and that the biologists themselves admit as much.
Primitive gene-vehicles evolved, in different ways and to different
extents, the typical characteristics of the ‘philosophical’ organism
because those that failed to do so lost out to the competition.
‘Organism’, we should conclude, is a vague concept, like
that of ‘material object’. A rock, a teacup, a railroad car, even a
skyscraper – these are all material objects. But what about a
mountain peak, a dish of oeufs a la neige, a plume of smoke, a
continent, a wireless network? These do not fit intuitive criteria
of physical massiveness and distinctiveness. Paradigm physical
objects can be moved, overturned, bumped up against or kicked
as well as pointed to and counted; the latter ‘semi-objects’ are
resistant to these forms of interaction. In this respect, they are
analogous to the forms of life that do not resemble the para-
digm objects mentioned by Ruse cited at the start of this essay.
Yet while the term ‘organism’ is typically contrasted with (mere)
‘material object’, there is an important difference between them.
The intensive investigation and exploration of the inner consti-
tution of one type of physical object – say a rock – will not
generalize in such a way as to assist in the understanding of a
teacup or a skyscraper, let alone a plume of smoke. But learning
about the inner constitution of sponges, flies, and other organ-
isms does provide insight into the basic processes of life taking
place in other organisms. This is because of the common origins
of life, its relatedness as branches of an evolutionary tree as-
sumed to be unique, and shaped by development in a very large
but still unique terrestrial environment.
The philosophical resonance of the notion is further appar-
ent from the moral value attached to it. We believe that organ-
isms are important, so that when the concept is said to be unim-
portant for natural science, the value of the things the term refers
Articles The Concept of Organism in the Philosophy of Biology 35

to is seen as disparaged. Titles like A Feeling for the Organism40


express moral concern for organisms and suggest that organisms
must be approached and appraised feelingly and that this feature
of life – that it is sensitive – is important. The conviction that
organisms, their powers, and their vulnerability are not sufficient-
ly appreciated, not only the belief that there will be explanatory
gaps in biology unless we allow full ontological status to organ-
isms, is perhaps behind the drive to save the concept of the
organism, as though saving the concept were a precondition of
saving the endangered things themselves.
There is accordingly a noticeable oversight in Dawkins’s
reconstruction of the organism as a walled – off, immunologi-
cal self-recognizer, other-rejector, and behavioral unit. He
forgot to mention that the organism does not only move its
limbs (when it has limbs) in a co-ordinated fashion – but sens-
es and reacts (when it has senses) as a unit. His language is, to
the contrary, consistently tough-minded. Asking why we do
not perceive the diminutive selfish replicators who cause nearly
everything of interest (including the weather!) to happen out-
side of astronomical and some geological and oceanographic
phenomena, he reminds us that the replicators have built
armor for themselves. They are encased in ‘gigantic lumbering
robots,’ sealed off from the outside world, yet manipulated by
remote control41. This imagery does not fit with modern eco-
logical sensibilities as well as the tender-minded depiction of
the autopoetic organism, acting, living, and preserving its
being. The deficiency is remediable, not by denying the projec-
tive component of the concept of the organism but by ac-
knowledging it, and by reminding ourselves that design con-
straints and ecological affordances render the vehicles not only
lumbering and well-armored, but in many instances, twitchy,
soft-bodied, swift, and graceful.

40 E. FOX-KELLER, A Feeling for the Organism: The Life and Work of Barbara
McClintock, W.H. Freeman, Boston 1984.
41 R. DAWKINS, Selfish Gene, cit., p. 19.
36 Catherine Wilson Articles

5. Conclusions

I hope to have supplied some answers to the question why


philosophers, writing out of traditions of social and political
philosophy in which notions of individuality, responsibility,
and integrity, and inviolability, as well as moral concern for
life, should believe that the concept of ‘the organism’ needs
and deserves rescue, and why practicing biologists are cavalier
about it. In Dawkins’s ‘dim vision,’ it is the invisible replicator
forces that possess robust reality and countable individuality.
Their vehicles are temporary, poorly-defined, fragile construc-
tions, identified by their phenotypical characters. These charac-
ters include behavior, both initiated and provoked, and all the
detached or detachable structures and products such as shells,
odors, nests, come-hither glances, dances, and seeds, that are
produced, as well as the ‘body’ in the more familiar sense of a
plant or animal or microorganism.
Questions about the relationship of reality as described by
the natural sciences to the categories of commonsense experi-
ence inevitably recur, as philosophers monitor and react to the
preconception-violating data of the special sciences. Our con-
cept of the self and so of the organism is shaped by the need of
human beings to initiate action, to react appropriately to threats
and provocations, and to experience and attribute responsibility,
as well as by the need to maintain a body in good condition.
Bodies like ours with sizable brains built by replicators for their
own persistence and reproduction will need the capacity to
represent and steer this machine around the environment. So
our core beliefs about the unity and identity of things and even
ourselves comes ‘from beneath,’ from those things inhabiting
the latent level that are not at all like us. This realization can be
disturbing, but, short of some form of mystical conversion, it is
likely impossible for us to think of the organism, and so of
ourselves, in any other way but the old-fashioned way, drawn
from the philosophical tradition. In that case, Dawkins’s ‘dim
vision’ will always seem a sort of revisionary metaphysics, a
form of theoretical conceptualization that can change how we
analyze specific events and occurrences, but never render the
Articles The Concept of Organism in the Philosophy of Biology 37

vernacular concept of the organism irrelevant unless, as noted


earlier, we cease to be organisms, phenotypes built by genotypes,
and the reflective automata that succeed us find nothing in the
rest of nature with which they can identify.
HOLISM, ORGANICISM
AND THE RISK OF BIOCHAUVINISM

by Charles T. Wolfe

Abstract. In this essay I seek to critically evaluate some forms of holism


and organicism in biological thought, as a more deflationary echo to Gilbert
and Sarkar’s reflection on the need for an ‘umbrella’ concept to convey the
new vitality of holistic concepts in biology (Gilbert and Sarkar 2000). Given
that some recent discussions in theoretical biology call for an organism concept
(from Moreno and Mossio’s work on organization to Kirschner et al.’s
research paper in Cell, 2000, building on chemistry to articulate what they
called “molecular vitalism,” studying the “vitalistic” properties of molecular,
cellular, and organismal function, and Pepper and Herron’s suggestion in
their 2008 paper that organisms define a category that evolutionary biology
cannot do without), the question, what concept of organicism are they calling
for? To what extent are such claims philosophically committed to a non-
naturalistic concept of organism as organizing centre, as a foundational rather
than heuristic concept – or possibly a “biochauvinism,” to use Di Paolo’s
term (Di Paolo 2009)? My aim in this paper is to conceptually clarify the
forms of holism and organicism that are involved in these cases (and I
acknowledge that the study of early 20th-century holisms [Peterson 2010]
indicates that not all of them were in fact ‘organicist’ or ‘biologistic’). I suggest
that contemporary holists are still potentially beholden to a certain kind of
vitalism or “biochauvinism”; but that when they reduce their claims to mere
heuristics, conversely, they risk losing sight of a certain kind of organizational
“thickness”, a “vital materiality” (Wheeler 2010) which is characteristic of
biological systems (Bechtel 2007). And I ask if it is possible to articulate a
concept of biological holism or organicism which is neither an empirical
‘biochauvinism’ nor a metaphysical ‘vitalism’?

I. A question sometimes asked in recent reflections at the in-


tersection of the philosophy of biology and so-called ‘historical
epistemology’ is, do organisms have an ontological status? Now,
we should note immediately that there are at least two very differ-
ent ways to answer this question in the affirmative. One is more
Verifiche XLIII (1-4), 2014, pp. 39-57.
40 Charles T. Wolfe Articles

philosophical, and seeks to provide criteria, e.g. of individuals


which organisms instantiate particularly well; from Aristotle to
Hegel and Jonas, this can be termed a strong holist option, but let
us again stress, in a philosophical context, for as we shall see, it is
possible to defend a holist claim for the ontological uniqueness of
organisms, on a biological basis as well (to be clear, I do not have
in mind a distinction between ‘metaphysical’ and ‘experimental’
here). A philosophical defense of strong holism as regards biolog-
ical individuals will tend to be more stipulative, appealing for
instance to the nature of whole/part relations and insisting that in
organisms, there is a special type of interrelation between parts.
In contrast, strong holism of the biological sort will typically
appeal to the empirical criteria defining actual organisms – but
which criteria?, would be the immediate retort, pointing to the
lack of any unifying concept for the diversity of actual organisms,
from drosophila to the Portuguese man-o-war1.
There are also diverse answers of the negative variety, which
I will be less interested in here. They include a kind of absolute
physicalism, which is more of an intellectual vue de l’esprit than a
position actually encountered amongst most working natural
philosophers: on this view, what is real is the entities specified in
fundamental physics, and since organisms do not belong to this
class of entities (however much some former physicists struggled
in the 1960s to find ‘organismic laws’, like Walter Elsasser2), they
can have nothing other than a ‘folk’, ‘cultural’ or perhaps ‘psycho-
logical’ status. More common is a general dismissal of the very
question, in the favor of a kind of evolving, instrumentalist con-
sensus, according to which there is no point in debating the
ultimate reality or ontological status of organisms, since what
matters is the theoretical terms science uses in its work, and these
constantly evolve (thus ‘organism’ gave way to ‘gene’, but even
‘gene’ now seems like a construct, since it can no longer be

1 J. PEPPER, M HERRON, Does biology need an organism concept? «Biological Re-


views», 83, 2008, pp. 621–627.
2 W.M. ELSASSER, Quanta and the concept of organismic law, «Journal of Theoretical

Biology», 1 (1), 1961, pp. 27-58 ; Reflections on a Theory of Organisms. Holism in


Biology, Johns Hopkins University Press, Baltimore, 1998.
Articles The Risk of Biochauvinism 41

thought of as «an inherently stable, discrete stretch of DNA that


encodes information for producing a protein», which is copied
faithfully before being passed on; rather, we now know that
«stability lies in the system as a whole, not in the gene»3).
Closer to my general interests, although I will not explore
this option here, is a more ‘historicized’ approach, in which the
reality of organisms is neither bluntly asserted – as is predomi-
nant in theoretical biology and most Continental ‘biophilosophy’4
– or denied, in favor of atoms, molecules, or selfish genes. In-
stead, it is examined in a diversity of historical régimes, contexts,
sometimes with an increased focus on the nature of the experi-
mental scientific apparatus involved. For instance, Hans-Jörg
Rheinberger describes experimental systems while emphasizing a
cognitive, constructed level of the effectivity of scientific practic-
es: experimental systems do not just inhabit a world of ‘things’
but also, what he calls «graphematic spaces», capable of «differen-
tial reproduction»5. Thus for our purposes, the reality of organ-
isms for an embryologist in the school of Wilhelm Roux is not
the same, e.g., as that experienced by an eighteenth-century
Montpellier medical vitalist insisting on the difference between
medical mechanistic models of the body and the ‘animal econo-
3 E. JABLONKA, M.J. LAMB, Evolution in Four Dimensions: Genetic, Epigenetic,
Behavioral and Symbolic Variation in the History of Life, MIT Press, Cambridge,
Mass., 2005, p. 7.
4 ‘Biophilosophy’ is a term that was popular in the post-war decades in French

‘épistémologie’ (and to some extent also in the German versions, where


authors like Kant and/or Hegel played a greater role, e.g. Gerhard Vollmer). It
was often presented as not being guilty of the reductionist excesses of analyti-
cally oriented ‘philosophy of biology’. My interest here is not to choose be-
tween the two but to judge the claims of certain forms of biological holism.
On the emergence of philosophy of biology as a discipline in contrast to
‘biophilosophy’ see J. GAYON, La philosophie et la biologie, in Encyclopédie
philosophique universelle, vol. IV, ed. J.-F. Mattéi, PUF, Paris, 1998, pp. 2152-
2171, which he has reprised with the more specific focus of ‘vitalism’ in his
Vitalisme et philosophie de la biologie, in Repenser le vitalisme, ed. P. Nouvel,
PUF, Paris, 2011, pp. 15-31.
5 H.J. RHEINBERGER, Experimental Systems – Graphematic Spaces, in T. LENOIR,

(ed.), Inscribing science: scientific texts and the materiality of communication, Stanford
University Press, Stanford, 1998, p. 287. Thanks to Claudia Manta for suggest-
ing this reference.
42 Charles T. Wolfe Articles

my’ (a term often used in that period to describe the functionally


interdependent, physiological system of the body, as distinct from
the mere machine). Curiously, both friends and foes of ho-
lism/organicism6 ignore the potential resources the historical
view would afford them, being content instead to assert or deny
an ontological state of affairs.
Here I will not pursue this more historical route, as I shall be
concerned with some varieties of biological holism, mainly in the
twentieth and twenty-first centuries, which I will seek to evaluate,
primarily in the form of a brief internal critique. To be clear, I am
not arguing for one sharply delimited ‘reductionist’ or ‘holist’
position, but rather, I am evaluating some of the tensions or
excessive commitments carried by various forms of holism, in
order to see what can be saved therein. My goal is to assess a
potential legitimacy of a certain kind of holism or organicism,
which, I shall suggest, needs to steer clear of a variety of wrong-
headed, anti-naturalistic or otherwise excessively falsifiable an-
swers to the initial question. To borrow an expression from
Daniel Dennett, the issue is: which varieties of organicism are ‘worth
wanting’? Since Scott Gilbert and Sahotra Sarkar’s reflection on
the need for an ‘umbrella’ or ‘organizing’ concept to convey the
new vitality of systemic or holistic concepts in biology7, seconded
by Manfred Laubichler’s paper proclaiming the return of the
‘organism’ as such an organizing concept8, some scholarly work
has been done which dispels earlier prejudices and gives us a
more useful, nuanced sense both of these concepts in biological
science and their possible pertinence today9.

6 I’ll refrain from using the potentially convenient two-letter acronym.


7 S. GILBERT, S. SARKAR, Embracing complexity: organicism for the 21st century,
«Developmental Dynamics», 219, 2000, pp. 1-9.
8 M. LAUBICHLER, The Organism is dead. Long live the organism! «Perspectives on

Science», 8 (3), 2000, pp. 286-315.


9 See e.g. P. HUNEMAN, C.T. WOLFE (eds.), The Concept of Organism: Historical,

Philosophical, Scientific Perspectives, special issue of History and Philosophy of the Life
Sciences, 32 (2-3), 2010 and T. CHEUNG, From the Organism of a Body to the Body of
an Organism: occurrence and meaning of the word ‘organism’ from the seventeenth to the
nineteenth centuries, «British Journal of the History of Science», 39, 2006, 319-
339, on the history and theory of organism.
Articles The Risk of Biochauvinism 43

In addition, there has been some sustained work on these


concepts in current biology. To name three recent examples, (1) in
theoretical biology, the effort to articulate a model of «organized
systems» and «organizational closure», in Alvaro Moreno and
Matteo Mossio’s research, alone or with collaborators10. A question
arising in reaction to this research is the extent to which philosoph-
ically it is committed to a non-naturalistic concept of organism as
organizing centre, as a foundational rather than heuristic concept –
or possibly a «biochauvinism», to use Di Paolo’s term11. He does
not define it, but we can imagine that biochauvinism would be the
present-day, naturalistic form of what was attacked by the Vienna
Circle and others as vitalism: no longer the claim that there are
immaterial vital forces which play a causal role in the functioning
of living beings, but simply the claim that living beings, as distinct
from tables, chairs and protons, are ontologically specific, not to
say special. This is not an essay devoted to the work of Moreno et
al. but I will state right away that I find their work especially free of
these more inflated metaphysical claims that are characteristic of
other, comparable theorists (including Francisco Varela, Robert
Rosen and Evan Thompson).
(2) In biochemistry, Kirschner et al.’s research paper in Cell
(Kirschner et al. 2000) on what they called «molecular vitalism»: they
suggested that, faced with the limitations of genomics, researchers
should investigate what the authors «whimsically» termed the «vital-
istic» properties of molecular, cellular, and organismal function: «the
organism has fashioned a very stable physiology and embryology
[…] It is this robustness that suggested ʻvital forcesʼ, and it is this
robustness that we wish ultimately to understand in terms of chem-
istry. We will have such an opportunity in this new century»12.

10 M. MOSSIO, A MORENO, Organisational closure in biological organisms, «History


and Philosophy of the Life Sciences», 32(2-3), 2010, pp. 269-288; A MORENO,
M. MOSSIO, Biological autonomy. A Philosophical and theoretical enquiry, Springer,
Dordrecht, forthcoming, and earlier, K. RUIZ-MIRAZO, A ETXEBERRIA, A
MORENO, J. IBÁÑEZ, Organisms and their place in biology, «Theory Bioscience»,
119, 2000, pp. 209-233.
11 E. DI PAOLO, Extended Life, «Topoi», 28, 2009, p. 10.
12 M. KIRSCHNER, J GERHART, T. MITCHISON, Molecular “Vitalism”, «Cell», 100,

2000, p. 87.
44 Charles T. Wolfe Articles

(3) In evolutionary biology, Pepper and Herron’s 2008 paper


suggests that organisms define a category that evolutionary biolo-
gy cannot do without, because it «relies heavily on the compara-
tive method, and effective comparison requires that we first
define a class of comparable entities», and they seek to articulate a
«robust and general operational definition» of organism, which
should distinguish between «which biological entities are organ-
isms and which are not», particularly from parts of organisms and
groups of organisms13.
The above examples indicate that different trends in biologi-
cal theory are articulating – or calling for – an organism concept.
However, only the first of these three cases directly contributes to
an explicitly holistic and organismic framework, which is the main
focus of this essay.
My aim is to conceptually clarify these forms of holism and
organicism. I acknowledge that the study of early twentieth-
century holisms indicates that not all of them were in fact ‘organ-
icist’ or ‘biologistic’. That is, some of the original forms of self-
described holism (I am not speaking of purported holists who
include such diverse figures as Spinoza, Hegel and Whitehead)
were primarily interested in systems as such, although they waver
on the issue, sometimes even seeking to derive principles from
biological systems for the understanding of other systems, as in
Walter Cannon’s suggestion that

It seems not impossible that the means employed by the more


highly evolved animals for preserving uniform and stable their
internal economy (i.e., for preserving homeostasis) may present
some general principles for the establishment, regulation and
control of steady states, that would be suggestive for other kinds
of organization—even social and industrial—which suffer from
distressing perturbations14.

13 J. PEPPER, M. HERRON, Does biology need an organism concept?, pp. 625, 621. For
more on the issue of ‘what units do biologists count?’ see the original article by
E. CLARKE, The Multiple Realizability of Biological Individuals, «The Journal of
Philosophy», CX (8), 2013, pp. 413-435.
14 W.B. CANNON, The Wisdom of the Body (1932), revised and expanded edition,

Norton, New York, 1963, p. 25 (thanks to Yelda Nasifoglu for help with this
reference).
Articles The Risk of Biochauvinism 45

What is holism? As mentioned above, it is intimately bound


up with systems theory, and with formally driven methods which
sought to give a new kind of coherence to biology, although this
‘systems’ interest was not always specifically biological. The
person often presented as the first theorist of holism, the South
African segregationist statesman (and specialist of Walt Whitman)
Jan Christiaan Smuts, as well as John von Neumann and Ludwig
von Bertalanffy later, wavers between defining holism as a total
systemic standpoint (with no particular reference to a special
status for living entities) and holism as an approach or model
which sheds particular light on embryology and how organisms
are not mere machines (with reference to teleology and the “his-
torical” or “learned” character of organisms).
Deliberately or inadvertently mirroring various other epi-
sodes in the history of mechanistic and organismic theories (in-
cluding the Leibniz-Stahl debate), these original holists also speci-
fy abstract terms on which ʻmerely mechanical aggregatesʼ are
different from genuine wholes, including chemical compounds,
and then specify that biological organisms are the exemplars of
ʻcreative wholesʼ, as Smuts calls them, i.e., wholes which create
structures different from their constituents or parts15. Here I will
not be concerned with holism as a general systems theory, but
rather with holism/organicism as a claim to define what is unique
in biological systems, i.e., a kind of biochauvinism.

II. Two critical remarks on two strangely apposite pitfalls of


holist-organicist theory:
(1) One notices a strange appeal of Kantianism for these
schools of thought, notably Varela’s16. There is something deeply

15 J.C. SMUTS, Holism and Evolution (1926), Sherman Oaks, CA, Sierra Sunrise
Books, 1999, pp. 140–141. The best general analysis and reconstruction of
holism and organicism in early twentieth-century biological thought is E.
PETERSONS’ dissertation, Finding Mind, Form, Organism, and Person in a Reduction-
ist Age, PhD, 2 vols., Program in History and Philosophy of Science, University
of Notre Dame, 2010.
16 Instances of such confusions can be found in A. WEBER, F.J. VARELA, Life

after Kant: Natural purposes and the autopoietic foundations of biological individuality.
46 Charles T. Wolfe Articles

flawed in a procedure which invokes the authority of the Kantian


‘projective’ approach to organisms17 in order to assert a set of
ontological specificities about organisms. First, because this is
precisely what the Kantian regulative ideal concept was designed
to avoid, in explicit contrast to what he would have called ‘ration-
al metaphysics’. To provide an empirical set of criteria for why
living beings are special and to claim that this supports or is
supported by a Kantian framework, is not a good idea if this
framework explicitly rejects the idea of giving empirical defini-
tions of organism, inasmuch as Kant’s organism concept is ex-
plicitly built around his notion of regulative ideal18. For Kant,
organism is a «reflective» construct rather than a «constitutive»
feature of reality, and reflective judgments are «incapable of
justifying any objective assertions»19.
Second, because it is not clear in any case why it counts as an
argument against ‘mechanistic science’ or ‘reductionism’ to simply
produce a list of key definitory features of Life, most classically,
self-preservation, self-reproduction, self-reparation, and self-
regulation20, or in a more updated form, «reproduction, life-
cycles, genetics, sex, developmental bottlenecks, germ-soma
separation, policing mechanisms, spatial boundaries or contiguity,
immune response, fitness maximization, cooperation and/or
conflict, codispersal, adaptations, metabolic autonomy, and func-
tional integration»21 (but precisely, the list is neither absolute nor
stable, which is why many thinkers, including prominent biolo-

«Phenomenology and the Cognitive Sciences», 1, 2002, pp. 97-125; S.


KAUFFMAN, G. LONGO, No Law Entails The Evolution Of The Bio-
sphere, 2011, http://lifeboat.com/blog/2011/07/no-law-entails-the-evolution
-of-the-biosphere; G. LONGO, M. MONTÉVIL, S. KAUFMAN, No entailing laws,
but enablement in the evolution of the biosphere, 2012, http://arxiv.org/pdf/
1201.2069.pdf
17 P. HUNEMAN (ed.), Understanding Purpose: Collected Essays on Kant and the

Philosophy of Biology, University of Rochester Press, North American Kant


Society Publication Series, Rochester, 2007.
18 I. KANT, Critique of Judgment (1790), tr. W. Pluhar, Hackett, Indianapolis,

1987, § 73, p. 276.


19 I. KANT, Critique of Judgment, § 67, p. 259; § 73, p. 277.
20 Cf. J. SCHLANGER, Les métaphores de l’organisme, Vrin, Paris, 1971, p. 14.
21 E. CLARKE, The Multiple Realizability of Biological Individuals, cit., p. 415.
Articles The Risk of Biochauvinism 47

gists, have declared that «Life as such does not exist»22), including
because mechanistic explanations have shown themselves to be
far more flexible and capable of integrating inter-level relations
than the depiction of them as taking apart systems into compo-
nent parts and leaving them there like broken toys.
Another puzzling feature of these projects in theoretical bi-
ology that invoke Kant but primarily seek to defend versions of
self-organization, might be expressed as: why the need for phi-
losophy? Why run the risk of lapsing into a kind of naïve, indeed
precritical metaphysics of selfhood and interiority, or of anti-
materialism? How could a working natural scientist take on board
proclamations such as Thompson’s: «Life is not physical in the
standard materialist sense of purely external structure and func-
tion. Life realizes a kind of interiority, the interiority of selfhood
and sense-making. We accordingly need an expanded notion of
the physical to account for the organism or living being»23? This
opposition between materialism and interiority is reminiscent of
post-Kantian denunciations of attempts to find the ʻseat of the
soulʼ. Further, here Thompson is calling for an expanded notion
of the physical, but often these kinds of theories, as I discuss
below, end up explicitly recommending that we disregard materi-
ality altogether – ironically given the desire on the part of, e.g.,
developmental systems theorists to move away from a disembod-
ied understanding of biological information, towards a kind of
vital materiality. More generally, if theoretical biology wants
organizational concepts, seeks to steer our attention away from
the sirens of genetic reductionism and the informational gene,
why should this entail being anti-materialists?) And I’ve noted
elsewhere24 that I think the appeal of Moreno, Mossio et al. is that

22 A. SZENT-GYÖRGI, The Living State: With Observations on Cancer, Academic


Press, New York, 1972, p. 1 (thanks to Gil C. Santos for this reference).
23 E. THOMPSON, Mind in life: biology, phenomenology, and the sciences of mind, Har-

vard University Press, Cambridge, Mass., 2007, p. 238. Would Thompson also
endorse Hegel’s assertion that « The spatiality of the organism has no truth
whatsoever for the soul » (G.W.F. HEGEL, Philosophy of Nature, trans. A.V.
Miller, Oxford University Press, Oxford, 1970, § 248Z, p. 18)?
24 C.T. WOLFE, Do organisms have an ontological status?, «History and Philosophy

of the Life Sciences», 32 (2-3), 2010, pp. 195-232.


48 Charles T. Wolfe Articles

they dispense with the somewhat foundationalist metaphysical


appeals of these other forms of holism.
(2) the problem of seeking to justifying the autonomy and/or
ontological uniqueness of Life, organisms, biological agents in
terms of empirical criteria, of a laundry list of properties of Life (from
the classic, self-preservation, self-reproduction, etc. as mentioned
above, to organizational closure, autonomy and so on). 25 As I
wrote a few years ago, one thinker’s homeostasis will always end
up being another thinker’s homeostat26; or, one theorist’s dynam-
ic equilibrium of hearts, kidneys, termite mounds or chemotaxis
will always end up being another theorist’s dynamic equilibrium
of storms or traffic jams. Not that all attempts to understand
such systems have to fall prey to the problems of a foundational
or otherwise ontologized empirical set of core criteria for Life:
Moreno and Mossio do not, as noted, and speaking of termite
mounds, when Turner presents them as a case of «embodied
homeostasis»27, he avoids these pitfalls since he does not ontologize
the property. The interaction between individual termites, and
between the termites and the termite mound, is a form of organi-
zation rather than a special substance, a point made brilliantly by the
systems theorist von Bertalanffy:

Organisms exhibit the properties of life not because of some


special peculiarity of these compounds, but on account of the
heterogeneous system into which these compounds are articu-
lated. There is no “living substance” because the characteristic
of life is the organization of substances28.

I note in passing, because that is a topic for a different paper,

25 For a more detailed, and sophisticated attempt to delineate a set of features


of ‘biological autonomy’, see B. ROSSLENBROICH, On the Origin of Autonomy – A
new look at the major transitions in evolution, Springer, Dordrecht, 2014.
26 That is, one thinker’s purportedly core biological feature can be artificially

modeled by another. C.T. WOLFE, Do organisms have an ontological status?. Thanks


to Olivier Surel for suggesting I clarify this.
27 J.S. TURNER, The Tinkerer’s Accomplice: How Design Emerges From Life Itself,

Harvard University Press, Cambridge, Mass., 2007, p. 27.


28 L.V. BERTALLANFY, Modern theories of development, trans. J.H. Woodger,

Oxford University Press, H. Milford, London, 1933, p. 48.


Articles The Risk of Biochauvinism 49

that one worthwhile holist-organicist strategy that does not invoke


definitional properties of life (or barely), is Kurt Goldstein’s, in
his suggestion that rather than say what is unique about the bio-
logical, we look to the observer: to be an organism is to have a point
of view on organisms; one which produces intelligibility, which
reveals organisms as meaning-producing beings29. This approach
– which has Kantian antecedents – is further extended after
Goldstein, by Georges Canguilhem.
Either way, waving lists of properties is a poor way to coun-
ter the power of reductionism. The reductionist can always reply,
in Pierre Duhem’s words, that a body is «only provisionally simple;
it has remained undecomposed until now, but tomorrow may
yield to a new means of analysis»30.

III. Holism or organicism, if it seeks to ontologize or other-


wise absolutize features of life (digestion, proprioception, inten-
tionality …) can run into some categorical problems, that is,
category mistakes, but also the danger of making empirical claims
to have identified absolute traits of life – à la Elsasser’s organis-
mic laws, or Varela’s «autopoiesis is an explication of the auton-
omy of the living»31, or Thompson’s statement of what we might
call the vulgate of biological holism, «parts arise from the whole.
Part and whole co-emerge and mutually specify each other»32,
itself not really an argument – which may in the end go up in
smoke, when it emerges that scientists cannot agree on a defini-
tion of life, and thus find it more helpful to eliminate the concept.
If the concept of Life is rejected, as in Albert Szent-György’s
«Life as such does not exist» or better-known, the partly per-

29 J. STAROBINSKI, L’idée d’organisme, Centre de Documentation Universi-


taire/Collège philosophique, Paris, 1956. The field of biosemiotics tarries at
the limit between this flexible, non-literal idea of organisms as meaning-makers
or world-makers, and a more stubbornly literal program.
30 P. DUHEM Le mixte et la combinaison chimique. Essai sur l’évolution d’une idée

(1902); reprint, Fayard-Corpus, Paris 1985, p. 50.


31 F. VARELA, Autonomy and Autopoiesis, in G. Roth and H. Schwegler, eds., Self-

Organizing Systems, Campus Verlag, Frankfurt 1981, p. 14.


32 E. THOMPSON, Mind in life, cit., p. 38.
50 Charles T. Wolfe Articles

formative statement by François Jacob, «On n’interroge plus la


vie aujourd’hui dans les laboratoires»33, claims to have found its
essential characteristics lose their foundation.
Further, sometimes the (former and current) holist and or-
ganicist insistence on opposing the systems they lovingly study to
some apparently inferior opposed entity, whether matter itself,
machines, corpses or the dehumanized world of Scientific Revo-
lution science, starts to sound ideological rather than scientific or
philosophically exploratory, even when it is put in deceptively
plain terms, such as here in Alva Noë’s recent book Out of our
heads: «to do biology we need the resources to take up a non-
mechanistic attitude to the organism»34. What would the mecha-
nistic attitude be? That of Descartes (who wrote, «the preserva-
tion of health has always been the principal end of my studies»35)?
Of Claude Bernard (for whom «what distinguishes a living ma-
chine is not the nature of its physico-chemical properties, com-
plex as they may be, but rather the creation of the machine which
develops under our eyes in conditions proper to itself and accord-
ing to a definite idea which expresses the living being’s nature and
the very essence of life»36)? Of the contemporary ‘mechanist’
William Bechtel, for whom mechanism and organization stand in
a complex and fruitful interaction? Mechanism, Bechtel suggests,
can provide an adequate account of organization by «placing as
much emphasis on understanding the particular ways in which
biological mechanisms are organized as it has on discovering the

33 F. JACOB, La logique du vivant, Gallimard, Paris, 1970, p. 320. See the inte-
resting reflections on this issue in Michel Morange’s review essay M. MO-
RANGE Un retour du vitalisme?, Histoire de la recherche contemporaine 2(2),
2013, pp. 150-155).
34 A. NOË, Out of Our Heads; Why You Are Not Your Brain, and Other Lessons from

the Biology of Consciousness, Farrar, Straus and Giroux, New York, 2009, p. 41.
35 In a letter to the Marquess of Newcastle, October 1645, in R. DESCARTES,

Œuvres, eds. C. Adam and P. Tannery, 11 vols., Vrin, Paris 1964-1976, vol. IV,
p. 329; The Philosophical Writings of Descartes, vol. 3: The correspondence, trans. J.
Cottingham, R. Stoothoff, D. Murdoch and A. Kenny, Cambridge University
Press, Cambridge, 1985, p. 275.
36 C. BERNARD, An Introduction to the Study of Experimental Medicine (1865), Henry

Schuman, New York, 1927, p. 93.


Articles The Risk of Biochauvinism 51

component parts of the mechanisms and their operations»37. But


this does not reduce to some pure mechanist ontology. In fact,
the holist/organicist standpoint minimally «remind[s] mechanists
of the shortfalls of the mechanistic accounts on offer», for ideas
such as «negative feedback, self-organizing positive feedback, and
cyclic organization are critical to explaining the phenomena
exhibited by living organism»38.
I say sometimes, since biochauvinism, or some weaker form of
it, need not be so ideological or oppositional. When it is not, it
can be appealing as a stubbornly realist or materialist-friendly
emphasis on genuine biological constraints. Consider the (some-
what restrictive) opposition Andy Clark proposes, between two
arguments for embodiment (and here for ‘embodiment’ we could
substitute ‘biochauvinism’, ‘organismicity’, etc.):

One of those strands depicts the body as special, and the fine
details of a creature’s embodiment as a major constraint on the
nature of its mind: a kind of new-wave body-centrism. The oth-
er depicts the body as just one element in a kind of equal-
partners dance between brain, body and world, with the nature
of the mind fixed by the overall balance thus achieved: a kind of
extended functionalism (now with an even broader canvas for
multiple realizability than ever before)39.

I daresay most materialists except ones especially enamored


of pure physics would want to say that «the fine details of a crea-
ture’s embodiment» count as «a major constraint on the nature of
its mind»; yet these fine details by no means imply that «the body
is special» in any way, especially not that of an embodied or
enactivist phenomenology founding itself on a claim of a founda-
tional interiority – recall, again, Thompson’s «Life is not physical
in the standard materialist sense of purely external structure and
function. Life realizes a kind of interiority, the interiority of

37 W. BECHTEL, Biological mechanisms: Organized to maintain autonomy, in F.


BOOGERD, F.J. BRUGGEMAN, J.-H.S. HOFMEYER, AND H.V. WESTERHOFF
(eds.), Systems Biology: Philosophical Foundations, Elsevier, Amsterdam, 2007, p. 270.
38 Ivi, pp. 296-297.
39 A. CLARCK, Pressing the flesh: a tension in the study of the embodied embedded mind ?,

Philos. Phenomenol. Res., 76 (1), 2008, p. 56.


52 Charles T. Wolfe Articles

selfhood and sense-making», to which we can add Varela’s claim


in a late paper that his project is to «reintroduce the subject into
biology»40. Now, I am not rejecting as anti-naturalistic the latter
part of Thompson’s claim – namely, that organisms are engaged
in ‘sense-making’, an idea we could fill in either ‘biosemiotically’
(with Uexküll’s analysis of Umwelt, in which «even the simplest
living organism creates a set of preferential partitions of the
world, converting interactions with their surrounding media into
elementary norms or values, as we will explain more extensively
below. And here is where the nature of living systems as autono-
mous agents, as inventors of worlds with meaning, becomes
manifest»41) or in ‘constructivist’ terms, appealing to Kurt Gold-
stein’s idea that an organism, unlike a skeleton or a watch, is an
meaning-producing entity. But these attempts to integrate ‘sense-
making’ or ‘world-making’ into the province of biology have
nothing to do with appeals to «the interiority of selfhood» or a
«reintroduction of the subject into biology».
Selfhood or subjectivity is not a useful biological concept,
any more than the soul was a useful medical concept (thus in the
eighteenth century, the Montpellier vitalist physician Ménuret de
Chambaud could seek to ‘eliminate’ soul from medical discourse,
in an entry on «Death» in the Encyclopédie42). There is too much
reliance here on an appeal to interiority, as if it was an unchal-
lengeable philosophical concept. As Jean-Marie Schaeffer says,
«In phenomenology, the problematic of embodiment (corporéité) is
part of an approach that continues to accept the epistemic privi-
lege of consciousness’s self-investigation as axiomatic»43. From a
naturalistic standpoint,

there are no intrinsically subjective or perspectival facts that are


either the special objects of self-regarding attitudes or facts of
“what it is like”. There are only states of subjects that both

40 A. WEBER, F.J. VARELA, Life after Kant, cit., p. 117.


41 K. RUIZ-MIRAZO, A. MORENO, Autonomy in evolution: from minimal to complex
life, «Synthese», 185 (1), 2012, p. 28.
42 J-J. MENURET DE CHAMBAUD, Mort, «Encyclopédie», vol. X, Briasson, Paris,

1765, p. 718b.
43 J.-M. SCHAEFFER, La fin de l’exception humaine, Gallimard, Paris, 2007, p. 118.
Articles The Risk of Biochauvinism 53

function in a particularly intimate way within those subjects and


have the subjects themselves and their other states as inevitable
referents. And that is all there is to “subjectivity” 44.

No one prevents philosophers from meditating on the inner


life, subjectivity or interiority. Thinkers from Augustine to Paul
Ricoeur have done so brilliantly45. But again, not only is selfhood
or interiority not a recognized biological concept; it is unclear
how it could play even a heuristic role therein (as opposed, e.g.,
to that of organization). To say it more positively, even if one is
interested, not in living systems in general, but in ‘minded’ living
systems, it remains the case that «things have a cognitive life
because intelligence exists primarily as an enactive relation be-
tween and among people and things, not as a within-intracranial
representation»46. We are better off thinking in terms of relations,
systems and interaction than in terms of a substantival life-force,
selfhood or interiority. But focusing on relations and networks
should not mean leaving out the materiality: «system thinking does
not imply forgetting about the material mechanisms that are crucial
to trigger off a biological type of phenomenon/behavior; rather, it
means putting the emphasis on the interactive processes that make
it up, that is, on the dynamic organization in which biomolecules
(or, rather, their precursors) actually get integrated»47.

44 W.G. LYCAN, What is the ‘Subjectivity’ of the Mental? Philosophical Perspectives 4:


Action Theory and the Philosophy of Mind, 1990, p. 126.
45 For a critique, see C.T. WOLFE, Éléments pour une théorie matérialiste du soi, in F.

Pépin, ed., La Circulation entre les savoirs au siècle des Lumières, Hermann, Paris,
2011, pp. 123-149.
46 L. MALAFOURIS, C. RENFREW (eds.), The cognitive life of things: recasting the

boundaries of the mind, McDonald Institute for Archaeological Research Publica-


tions, Cambridge, 2008, Introduction, p. 4. One can also be more charitable to
Thompson’s enactivism, as he is not always unhappily insisting on a founda-
tional interiority (what some would call ‘internalism’); he also describes living
interiority as «compris[ing] the self-production of an inside that specifies an
outside to which that inside is constitutively and normatively related» (E.
THOMPSON, Mind in life, cit., p. 225).
47 K. RUIZ-MIRAZO, A. MORENO Basic Autonomy as a Fundamental Step in the

Synthesis of Life. Artificial Life, 10, 2004, p. 238.


54 Charles T. Wolfe Articles

IV. In sum, holism/organicism can succumb to two major


kinds of temptation: it can present itself as an empirical theory,
providing the key empirical characteristics of organisms, over and
against the rest of the natural world. Or, conversely, there is what
one might call the transcendental problem, in the sense of a
transcendentalizing of organism, or the flesh, or embodiment.
One can indeed be reluctant to conceive of the organism in
purely computational terms without invoking a mysticism of the
flesh (from Merleau-Ponty’s appeals to transubstantiation in
seeking to highlight what is unique in the sensation of an embod-
ied being48 to Thompson’s denial that «life is physical in the
materialist sense»). Here, a third problem emerges: the addiction
to formalisms at the expense of materiality.
Faced with certain holistic excesses, I am not, to repeat, ad-
vocating the extreme prudence of instrumentalism, in which the
moral when it comes to the material realization of biological
systems might be ‘handsome is as handsome does’. Contempo-
rary holists are often still potentially beholden to a certain kind of
vitalism or «biochauvinism»49; but when, out of extreme caution,
they reduce their claims to mere heuristics, conversely, they risk
losing sight of a certain kind of organizational «thickness», a «vital
materiality» in Wheeler’s terms50; he opposes vital to implementa-
tional materiality, without really fleshing out the distinction, so to
speak, but I assume it is similar to what an earlier, perhaps the
first theorist of vital materiality, Denis Diderot, expressed in a

48 «Just as the sacrament not only symbolizes … an operation of Grace, but is


also the real presence of God … in the same way the sensible has not only a
motor and vital significance but is nothing other than a way of being in the
world that our body takes over […] sensation is literally a communion» (M.
MERLEAU-PONTY, Phénoménologie de la perception, Gallimard, Paris, 1945, p. 245).
49 Again, this is not an essay in the historical epistemology of organisms, or a

suggestion in how to rethink our presuppositions about vitalism, e.g. by


revisiting its history. For some efforts in this direction, see C.T. WOLFE, From
substantival to functional vitalism and beyond, or from Stahlian animas to Canguilhemian
attitudes. Eidos, 14, 2011, pp. 212-235; S. NORMANDIN, C.T. WOLFE (eds.),
Vitalism and the scientific image in post-Enlightenment life science, 1800-2010, Springer,
Dordrecht, 2013.
50 M. WHEELER, Mind, things and materiality, in L. MALAFOURIS, C. RENFREW

(eds.), The cognitive life of things, pp. 29-37.


Articles The Risk of Biochauvinism 55

nice metaphor: «What a difference there is, between a sensing,


living watch and a golden, iron, silver or copper watch!»51.
If there was only implementational materiality, (i) we would
fail to grasp any meaningful difference between a living watch
and a silver or copper watch – and please note, this difference
need not be justified in terms of a vital force, an entelechy, an élan
vital, a centralizing soul or self, an irreducible intentionality or
first-personness52 – and (ii) we would lapse into a kind of lazy
multiple realizability position53 which disregards the material in
which a system is realized, of the sort classically stated by Varela
and later, Robert Rosen. Here is Varela:

We are thus saying that what defines a machine’s organization is


relations, and hence that the organization of a machine has no
connection with materiality, that is, with the properties of the
components that define them as physical entities. In the organiza-
tion of a machine, materiality is implied but does not enter per se.54

Using much the same language, Rosen describes his ap-


proach as «relational», which we might think of as congenial to
the study of specific biological forms of organization, but no: he
adds that the relational approach aims to «throw away the matter
and keep the underlying organization»55. This is like the central
dogma of functionalism, i.e., multiple realizability, by virtue of
which the materiality of systems is held to be irrelevant: as ex-
pressed classically by the arch-functionalist Hilary Putnam, «we
51 D. DIDEROT, Éléments de physiologie, in Œuvres complètes, dir. H. Dieckmann, J.
Proust, J. Varloot, Éditions Hermann, Paris, 1975-, XVII, p. 335.
52 Contra the Varela first-person business (e.g. F.J. VARELA, J. SHEAR, First-

person methodologies: why, when and how, Journal of Consciousness Studies, 6(2-3)
1999, pp. 1-14).
53 See the excellent history and analysis of multiple realizability arguments in J.

BICKLE, Multiple Realizability, in E. ZALTA (ed.), Stanford Encyclopedia of Philosophy,


http://plato.stanford.edu/entries/multiple-realizability/, 1998; updated 2013.
54 F.J. VARELA, Principles of Biological Autonomy, Elsevier North Holland, New

York, 1979, p. 9, cit. in W. BECHETEL, Biological mechanisms, cit., p. 294. To be


fair, in context Varela is not calling for multiple realizability, but it is not clear
why or how he is not.
55 R. ROSEN, Life Itself, Columbia University Press, New York, 1991, p. 119; cf.

«throwing away the physics and keeping the underlying organization» (p. 280).
56 Charles T. Wolfe Articles

could be made of Swiss cheese and it wouldn’t matter»56.


To conclude, there is something endearing about vitalist
claims especially when they don’t invoke or rely on a foundational
subjectivity, including because they don’t revert to explaining
biological systems (embodied agents) in fully systemic, formalized
terms. So is a small amount of biochauvinism acceptable, or even
desirable? If it means the effort to grasp the kind of vital material-
ity characteristic of biological systems – their organization, in the
language of Moreno, and also Bechtel, who, like Scott Turner,
invokes Claude Bernard as an important predecessor of this
view57. Recall that for Bernard, «In order to study the phenomena
pertaining to living beings and discover the laws that govern
them, it is not necessary to know the essence of life itself»58. We
could enlist Bernard in support of the view that an investigation
of vital materiality need not transcendentalize Life or organism.
But is this materiality «substance» or «organization»? It is not
quite a biochauvinism, or very weakly, because it does not posit a
vital substance, a foundational interiority or an absolute set of
criteria for Life: recall von Bertalanffy’s observation, «There is no
“living substance” because the characteristic of life is the organi-
zation of substances»59. However, as I’ve said earlier, a focus on
organization and relations need not imply that we sacrifice its
materiality in favour of ‘Swiss cheese’ multiple realizability. A
biochauvinism implies a degree of ontological commitment (what
I termed a «weakly ontological» view of organism in my earlier
reflection on the topic60). A tangibility, implicit in what I above

56 H. PUTNAM, Philosophy and our Mental Life, in Putnam, Mind, Language, and
Reality. Philosophical Papers, vol. 2, Cambridge University Press, Cambridge,
1975, p. 291.
57 J. S. TURNER, cit., chapter 2 (on Bernard machines); idem, Homeostasis and the

forgotten vitalist roots of adaptation, in S. NORMANDIN, C.T. WOLFE (eds.), Vitalism


and the scientific image, cit.
58 C. BERNARD, Histoire de l’expérimentation physiologique – l’art d’expérimenter sur les

êtres vivants, Revue des cours scientifiques de la France et de l’étranger, 6e


année, Germer Baillière, Paris, 1869, p. 194.
59 L.V. BERTALLANFY, Modern theories of development, trans. J.H. Woodger,

Oxford University Press, H. Milford, London, 1933, p. 48.


60 C.T. WOLFE, Do organisms have an ontological status?, cit.
Articles The Risk of Biochauvinism 57

pointed to – that most materialists can accept a degree of embod-


iment (a.k.a. biochauvinism), granting that facts about bodies can
act as ʻmajor constraintsʼ on facts about minds. Yet this tangibil-
ity, this embodiment appeal to no foundational interiority, no
special inner life.
My suggestion then is that it should be possible to articulate
a concept of biological holism or organicism (whether it is locat-
ed in systems biology, theoretical biology, evolutionary biology or
a philosophical reconstruction of several of these) which dispens-
es with the first-person obsession or the transubstantiation-
friendly invocation of an ontology of the body as corps propre,
although this is not the same as Clark’s, Wheeler’s or Malafouris’
attention to vital materiality. Such a view is neither an empirical,
laundry list ‘biochauvinism’ nor a metaphysical ‘vitalism’.

Acknowledgments. Earlier versions of this paper were presented at the


Embodiment and Adaptation Workshop, University of Pittsburgh, Center for
Philosophy of Science (March 2011) and ISHPSSB, Montpellier (July 2013).
Thanks in particular to Matteo Mossio for his comments, although he would
not endorse the contents of the paper.
ORGANISMICITY AND THE DECONSTRUCTION OF
THE ORGANISM: FROM SUBSTANTIAL FORMS TO
DEGREES OF COOPERATION, CLOSURE AND AGENCY

by Georg Toepfer

Abstract. Since antiquity living beings have been conceptualized as discrete


entities with clear-cut boundaries in space and time. However, the scientific
understanding of living beings as ‘organized bodies’ or ‘organisms’ has result-
ed in the difficulty of specifying living systems by one single criterion or a set of
clear-cut criteria. Moreover, all the criteria that were used for the identifica-
tion of organisms – especially functional closure, cooperation and agency –
admit of degrees. Therefore, organisms seem to be at an extreme point on an
entire spectrum of organized systems in nature. Given this situation, the
notion of ‘organismicity’ in the sense of a state variable, i.e., a property of
systems that comes in degrees, seems suitable in many contexts. It adequately
expresses the gradient nature of various aspects of the organism’s way of
being, in particular the following six: (1) the transition from organized
systems in nature to organisms, (2) the beginning and ending of an organism
in time, (3) the spatial limits of individual organisms in cases of clonal
growth, (4) the existence of various levels of organization within an organism,
(5) the often vague boundary between organism and environment, and (6) the
occasionally gradual transition from individual to community.

Organisms provide the best examples of integrated and func-


tionally closed systems in nature. However, these systems are not
easily specified by one single criterion or a set of clear-cut criteria.
Moreover, all the criteria we use for the identification of organ-
isms admit of degrees. Organisms seem to be at an extreme point
in a whole spectrum of organized systems in nature. Given this
situation, the notion of ‘organismicity’1 in the sense of a state
variable, i.e. a property of systems that comes in degrees, seems

1The term ʻorganismicityʼ appeared in the first half of the 20th century and
was casually used by various authors (see www.biological-concepts.com: s.v.
“organismicity”).
Verifiche XLIII (1-4), 2014, pp. 59-75.
60 Georg Toepfer Articles

to be more suitable, at least in some contexts, for describing the


phenomenon than the notion of ‘organism’, which denotes indi-
vidual systems themselves and not their type of causal structure.
Ironically, in the first century of its existence, from the end
of the 17th to the end of the 18th century, the very notion of ‘or-
ganism’ was used in the sense now associated with the state
variables ‘organization’ and ‘organismicity’: In the writings of
Georg Ernst Stahl, Gottfried Wilhelm Leibniz and Nehemiah
Grew around 1700, ‘organism’ was a principle of order in certain
types of systems, especially living systems. In a sense it was op-
posed to the concept of ‘mechanism’. Grew, for example, consid-
ered the ‘organism’ as the organizing principle of a living body; he
spoke of the «organism of a body» and stated that «every body
should have its Organism»2. This means, a living body is not an
organism but has an organism. Grew and other 18th-century
authors used the word ‘organism’ in much the same way we use
the expression ‘organization’. By speaking of ‘organismicity’
instead of ‘organisms’, we tend to return to this original meaning.
With ‘organismicity’ as a state variable we can express the multi-
tude of ways in which living beings are structurally ordered on
various levels of organization.
My account of this shift from ‘organism’ to ‘organismicity’
proceeds in five steps: I will first turn to the old understanding of
individual living beings as possessing substantial forms which has
been a well-established notion since antiquity. I will then identify
three central characteristics of individual living beings that were
important in developing the concept of ‘organism’, namely func-
tional closure, cooperation and agency. In each section, I will show how
all three terms come in degrees. The final part addresses the
consequences of this view of organisms as systems with gradually
varying properties.

2N. GREW, Cosmologia sacra, Rogers, Smith and Walford, London 1701, p. 24;
T. CHEUNG, From the Organism of a Body to the Body of an Organism: Occurrence and
Meaning of the Word ‘Organism’ from the Seventeenth to the Nineteenth Centuries,
«British Journal for the History of Science», 39, pp. 319-339.
Articles Organismicity and Deconstruction of the Organism 61

1. Organisms as Substantial Forms

Since antiquity organisms have been conceptualized as discrete


entities with clear-cut boundaries in space and time. For Aristotle,
living beings were paradigmatic cases of substantial forms. They
were characterized by very general functional capacities such as
nutrition, growth and reproduction. These capacities, in turn, only occur
in systems of a certain type, namely in living beings, and these beings
were individuated by a specific principle, the soul3.
A separate soul was attributed to every living being; the soul
was the ontological and explanatory principle that unifies all
living beings into a distinct class of entities. Although in Plato’s
and Aristotle’s conception the soul had hierarchical parts and
only one part, the lowest one, was common to all living beings,
the question whether or not to attribute a soul to an individual
being was itself unambiguous.
In the concept of substantial form, the term ‘form’ was relat-
ed to the specific form or design of an organism. But as it was
connected to physiological processes itself, it was thought of as a
dynamic principle “which causes not only the unity, identity and
being of the living body, but also its biological processes”4. In this
respect, the Aristotelian soul was a principle that is similar to the
modern concept of ‘organization’. Following Martha Nussbaum’s
book about De motu animalium, modern notions of ‘dynamic
organization’ or ‘functional organization’ are often used as trans-
lations for the Aristotelian concept of a soul5. As Diana Quaran-
totto has indicated, this translation seems to be fitting in several
respects: (1) Like the soul, the organization is not a body in itself
but something belonging to a body. (2) Soul and organization do
not exist independently of the living body. (3) Like the soul, the
organization is considered as the principle of unity and identity of
the body. And (4) soul and organization are both principles of

3ARISTOTLE, De anima: 415a 21-b 2.


4 D. QUARANTOTTO, Aristotle on the Soul as a Principle of Biological Unity, in S.
FÖLLINGER (ed.), Was ist ,Leben’? Aristotelesʼ Anschauungen zur Entstehung und
Funktionsweise von Leben, Steiner, Stuttgart 2010, pp. 35-53.
5 M. NUSSBAUM, Aristotle on Teleological Explanation, in Aristotle’s De Motu

Animalium, Princeton University Press, Princeton 1978, pp. 59-106.


62 Georg Toepfer Articles

self-movement in dynamic systems.


But despite these parallels there is one important distinction:
Whereas ‘organization’ is a concept that comes in degrees, a body
either has a soul or does not. For Aristotle it seemed reasonable
to postulate a principle like the soul because living beings ap-
peared to him as a clear-cut class of entities. To be sure, Aristotle
thought it possible for many kinds of organisms to come into
being by spontaneous generation. He stated in this context that
even the earth and water were filled with spiritual heat and soul6,
but nevertheless, according to Aristotle, the actualization of the
soul takes place only in living beings. The soul, therefore, was
without question a principle of life.
This situation changed with the identification of living beings
with ‘organized systems’ or ‘organisms’. This identification took
place in the context of mechanistic accounts of living beings
starting in the middle of the 17th century. In his Passions of the Soul
(1649) René Descartes explicitly stated that the functions of life
depend on the arrangement or disposition of the organs («la
disposition de ses organes»7. In 1662 the botanist Joachim
Jungius replaced ‘soul’ with ‘organization’ as the central principle
for the explanation of life functions. With the explanatory princi-
ple of ‘organization’ the soul becomes superfluous: «vero
organisatio sola sufficiat», as Jungius put it8, or «L’organisation et
la vie, voilà l’âme», as Diderot wrote a century later9.
The situation has not changed since: We see life as a conse-
quence of the organization of matter. From the 18th century on

6 ARISTOTLE, De generatione animalium, 762a 18; W. KULLMANN, Übergänge


zwischen Unbeseeltheit und Leben bei Aristoteles. In: S. FÖLLINGER, (hg.), Was ist
,Leben’? Aristotelesʼ Anschauungen zur Entstehung und Funktionsweise von Leben,
Steiner, Stuttgart 2010, p. 134.
7 R. DESCARTES, Les passions de l’ame (1649), in C. ADAM, & P. TANNERY (eds.),

Œuvres de Descartes, vol. XI, Vrin, Paris 1986, p. 351.


8 J. JUNGIUS, De vita plantarum, in Doxoscopiae physicae minores, Part 2, Section 3,

Fragment 5, Naumann, Hamburg 1662, p. I.


9 D. DIDEROT, Éléments de physiologie (1778), ed. P. Quintili, Champion, Paris

2004, p. 358; see also: J.T. NEEDHAM, Nouvelles observations microscopiques, avec des
découvertes intéressantes sur la composition et la décomposition des corps organises, Ganeau,
Paris 1750, p. 375.
Articles Organismicity and Deconstruction of the Organism 63

‘organization’ has been the central concept in definitions of ‘life’,


as explicitly stated in Voltaire’s simple characterization: «Life is
organization with the capacity of feeling»10; or in Girtanner’s and
Kant’s identification of the meanings of the words ‘living’ and
‘organized’: «Les mots organisé & vivant sont, selon moi, des
synonimes»11; «Daß ein organischer Körper belebt ist, ist ein
identischer Satz»12.
In the 19th and 20th centuries we see the equation of life with
organization in the development of theories of self-organization
(«Leben ist Organisation, Selbstorganisation»13) or in systems-
theoretical accounts like that of Ludwig von Bertalanffy («Die
Besonderheit des Lebens beruht […] auf Organisiertheit»14).
For many authors, ‘organization’ seemed to be an appropri-
ate concept for analyzing ‘life’ because it allowed combining the
structural heterogeneity of a system with functional homogeneity. De-
scribed as ‘organized’, living beings could be seen as structurally
complex and teleologically ordered at the same time: Organized
systems are composed of many different parts and particles, but
the composite whole of these diverse interacting parts strives for
a very restricted set of goals, self-preservation and reproduction being
the two most important. The concept of ‘organization’, therefore,
makes it possible to give an account of living beings on a mecha-
nistic basis as intrinsic natural unities or wholes. With its focus on
interacting parts, the concept is mechanistic, but in stressing the
emergence of complex functions at the level of their interaction it
is holistic.
The main mechanism of organization that guarantees this
double perspective is the mutual dependence of the parts. This is a

10 VOLTAIRE, Vie, in Questions sur l’encyclopédie, vol. 9. Paris 1772, pp. 55-58.
11 C. GIRTANNER, Mémoires sur l’irritabilité, considérée comme principe de vie dans la
nature organisée, in Observations sur la physique, sur l’histoire naturelle et sur les arts, 37,
Ruault, Paris 1790, p. 150.
12 I. KANT, Opus postumum, in A. BUCHENAU (ed.), Kant’s Opus postumum, Vol. I

(Akademie Ausgabe, vol. XXI), Reimer, Berlin 1936-38, p. 66.


13 K. FISCHER, System der Logik und Metaphysik oder Wissenschaftslehre, 2nd ed.,

Bassermann, Heidelberg 1865, p. 524.


14 L. BERTALANFFY, Kritische Theorie der Formbildung, Borntraeger, Berlin 1928,

pp. 68-69.
64 Georg Toepfer Articles

concept with ontological implications; it is not only causal as


‘interaction’ is. A description of a system in terms of mutual
dependence is holistic because it stresses the ontological dependence
of each part on the cooperation with the other parts. Since the
mid-17th century mutual dependence has gained a primary position
in the analysis of living systems. For example, in 1644 Kenelm
Digby compared living beings to machines based on things he saw
on his trip to Spain. He identified the mutual dependence between
the parts as the central common organizing principle between
machines and living beings. As he put it, in these systems «the one
[part is] not being able to subsist without the other, from whom he
deriveth what is needefull for him; and again being so usefull unto
that other and having its action and motion so fitting and necessary
for it, as without it that other can not be»15.
Since that time the principle of mutual dependence has be-
come central to most definitions of living beings as organisms. A
sampling from four short definitions in four different languages
attests to this: In 1675, Nicolas Malebranche defined an organic
body («corps organisé») as a whole consisting of infinite parts that
mutually depend on each other («qui dépendent mutuellement les
unes des autres»16). Sixteen years later John Ray said of the parts
of an organized being that «they do all Friendly conspire, all help
and assist mutually one the other, all concur in one general End
and Design, the good and preservation of the whole»17. In the
physiological context, Herman Boerhaave defined an organic
body («corpus Organicum») by the interdependence of its parts
(«harum partium actiones ab invicem dependent»18. And, finally,
the most famous definition is Immanuel Kant’s from 1790,
which was probably inspired by Boerhaave, whose students

15 K. DIGBY, Two Treatises in the one of which the Nature of Bodies, in the other, the
Nature of Man’s Soule is Looked into, Blaizot, Paris 1644, p. 205; T. CHEUNG, Res
vivens. Agentenmodelle organischer Ordnung 1600-1800, Rombach, Freiburg im
Breisgau 2008, p. 25.
16 N. MALEBRANCHE, De la recherche de la vérité (1675), vol. 2, David, Paris 1721,

p. 57.
17 J. RAY, The Wisdom of God Manifested in the Works of the Creation (1691), Smith,

London 1961, p. 216.


18 H. BOERHAAVE, Historia plantarum, Gonzaga, Rome 1727, p. 3.
Articles Organismicity and Deconstruction of the Organism 65

were teaching physiology in Konigsberg at the time. According


to Kant’s definition, «in a thing as a natural purpose» the parts
combine in the unity of a whole («verbinden sich zur Einheit
eines Ganzen») insofar as they are reciprocally cause and effect
of each other’s form19.
In all these definitions it is the mutual dependence of the
parts that is responsible for the system’s unity and closure: the
common principle is based upon the unity of a complex by the inter-
dependence of its parts. In this way it was possible to reconstruct
living beings as intrinsic units of nature on a mechanistic basis.
Accordingly, the function of the soul as lending a natural body
complex capacities and unity is fulfilled by the principle of inter-
action and interdependence.
But replacing the soul by ‘organization’ is the first and most
important step in the dissolution of the phenomenon of life into
a category with many indistinct, blurred edges. In the next section
I will discuss this dissolution in three steps by distinguishing
between closure, cooperation and agency as aspects of systems unified
by the interdependence of their parts.

2. Closure in Degrees

The appearance of living beings as closed and isolated enti-


ties was one of the strongest arguments for their consideration as
substantial forms, as intrinsic natural units. Yet, closure in itself is
a complex phenomenon, and there are different kinds of closure.
The first important distinction is between structural (morphologi-
cal) and functional (physiological) closure. Structural closure, the
connection of the parts in space, is something living beings have
in common with inorganic bodies such as stones. In a simple
classification there are two types of morphological closure: contigu-
ity, i.e., the spatial holding together of parts, and delimitation, i.e.,
the isolation of the system from its environment by structures
such as membranes. In contrast to this, functional closure is a feature

19 I. KANT, Kritik der Urtheilskraft, in Kant’s gesammelte Schriften, Königlich


Preußische Akademie der Wissenschaften, Bd. V., Reimer, Berlin 1913, p. 373.
66 Georg Toepfer Articles

of living beings that most inorganic bodies do not have. It con-


sists in the mutual dependence of the parts in a system. In its
basic form it is a cycle of processes in which each process type
influences another process type in the system such that the effect
of each process type feeds back on itself. The simple scheme of a
causal cycle can further be complicated by one part influencing
and supporting more than one other part, i.e., by the integration of
the system. In contrast to cyclicity, which is an all-or-nothing
phenomenon, integration comes in degrees and could be defined
as a quantitative variable. The same relation holds true for conti-
guity and delimitation: contiguity is an absolute phenomenon;
delimitation is gradual and never complete.
In what follows I shall focus on physiological closure as it is
the characteristic type of closure in organisms. Physiological
closure, or functional closure, is in itself not a very restrictive criteri-
on because it can be applied to very different kinds of systems,
not only to living systems. And even within the class of living
systems it is a liberal criterion that allows for the identification of
systems at various levels. In purely functional terms there is not
only one organismic level. Imagine, for instance, a spider in its
web. In a sense there are two organisms involved in this case: one
is the spider as a morphologically and physiologically closed
system, and the other is the spider together with its web. In the
second case the web is part of an organism because it was pro-
duced by the metabolism of the spider and it feeds back on this
metabolism by contributing to the spider’s nutrition. The criteri-
on of interdependence alone is ambiguous here, as it justifies the
identification of two overlapping functionally closed systems.
Additionally, inorganic systems may exist that fulfill the con-
ditions of interdependent parts and processes. Hence, the mutual
dependence of parts may be a central aspect of natural organisms
but it is certainly not a sufficient criterion for their identification.
Additional criteria for specifying the causal structure of living
organisms are necessary.
The second point about the existence of systems of mutually
dependent processes in the inorganic world often involves a dis-
cussion of the water cycle. It consists of three mutually dependent
processes: evaporation, condensation and precipitation. Due to the cycli-
Articles Organismicity and Deconstruction of the Organism 67

cal connection with other processes, each instance of precipitation


feeds back onto future instances. Consequently, in a systems-
theoretical account of function, it would be legitimate to assign a
function to this and the other sub-processes of the water cycle.
With respect to causal cyclicity and mutual dependence, the
water cycle and a living organism are parallel cases. But there are,
of course, differences in their causal structure. One important
difference is that living organisms have a coherent body. They are
embodied structures. This embodiment makes an important
difference because the body functions as a self-constructed
boundary condition for physical laws. In the early 1970s Howard
Pattee emphasized this aspect of living systems as one of their
most characteristic features. According to Pattee, «Life is distin-
guished from inanimate matter by the co-ordination of its con-
straints»20. The self-imposed boundary conditions or «constraints»
of the system realized in the organic body control the integrative
behavior of the system. This control of physical laws and com-
ponent parts of the system forms the basis for the freedom and
adaptability of the whole organism21.
In later accounts, Alvaro Moreno, Jon Umerez and Julio
Fernández elaborated on the concept of constraints and used it
for a definition of life. In their view, living systems are a particu-
lar type of chemical network with a circular organization main-
tained by the system itself. In contrast to inorganic networks with
a circular organization like the water cycle, which are governed by
the general laws of nature (such as the laws of evaporation and
condensation of water), living systems are additionally subject to

20 H.H. PATTEE, Physcial theories of biological co-ordination, «Quarterly Review of


Biophysics», 4, 1971, pp. 255-276, p. 273.
21 Ivi, p. 256. An early equivalent to the concept of ‘constraints’ in this context

is «physical topography». It appeared in the classical debate between Wolfgang


Köhler and Hans Driesch in the 1920s about the distinction between living
organisms and physical local patterns (Gestalten). According to Köhler,
reference to the “physical topography” of an organism is sufficient to account
for its coordinated activities. See: W. KÖHLER, Die physischen Gestalten in Ruhe
und im stationären Zustand, Vieweg, Braunschweig 1920, p. 79; H. DRIESCH,
“Physische Gestalten” und Organismen, «Annalen der Philosophie und
philosophischen Kritik», 5, 1926, pp. 1-11, p. 5.
68 Georg Toepfer Articles

«local constraints created by the localized action of certain com-


ponents, in such a way that the stability of the organization forms
is linked to individual components»22. In this sense, biological
organisms «realize a specific kind of causal regime […], a distinct
level of causation, operating in addition to physical laws, generat-
ed by the action of material structures acting as constraints»23.
Moreno and his colleagues argue that in the case of the water
cycle there is no organizational closure as there is no distinct
body produced by the system that exerts a constraining influence
on its physical laws and processes. I would prefer to say that due
to the interdependence of the processes there is organizational
closure in this case as well (and, consequently, it is legitimate to
assign functional roles), but the water cycle system does not meet
the additional criterion of having a body and lacks the particular
causal regime associated therewith due to the local constraints
imposed on the components by this structure. This is true for
other biological systems that are organizations as well. Ecosystems,
for example, do not have one coherent body, but there are clearly
causal roles in these systems. It is also common practice in biolo-
gy to assign functions to the bearers of ecological roles, as in
‘producers’ or ‘reducers’ of organic matter.
In my view, having a body is an important feature of living
organisms because it turns these organizations from merely
functionally into structurally (morphologically) closed systems. In
contrast to the global character of the water cycle, living organ-
isms exist in discrete pieces of matter. Many of these pieces can
exist side by side. Since they can reproduce and show variation in
their reproduction, they are objects of natural selection. In the
process of evolution, natural selection can lead to an increase in
complexity, integration and stability of these systems. In the
history of the earth, the escalation of functional complexity has
only taken place in living systems and has been closely connected
to increases in structural complexity as an ever growing constrain-

22 A. MORENO, J. UMEREZ, J. & J. FERNANDEZ, Definition of life and the research


program in artificial life, «Ludus Vitalis», 2, pp. 15-33, p. 17.
23 M. MOSSIO, M. & A. MORENO, Organisational closure in biological organisms,

«History and Philosophy of the Life Sciences», 32, 2010, pp. 269-288, p. 269.
Articles Organismicity and Deconstruction of the Organism 69

ing factor. The final results of this process are the immensely
complex systems of living beings we encounter today.
This increase in complexity and integration is not possible in
the case of the water cycle because it is a single global system that
is not subject to natural selection. But in its causal pattern of
interdependent processes the water cycle is nevertheless similar to
living organisms. With its globally dispersed character it is located
at one end of the continuum of embodied structures, while or-
ganisms are at the other end with maximal embodiment.

3. Cooperation in Degrees

Cooperation is a phenomenon that can take place within and


between organized systems. In a general sense, every organization
is a co-operation in itself insofar as each of its mutually depend-
ent parts contributes to the maintenance of the other parts and
the system as a whole.
In an interpretation of Aristotle’s conception of the soul,
Miller and Miller gave the following definition of a cooperative
organization:

By «cooperative» organization we mean a configuration of rela-


tions in which the subfunctional components mutually enable
each other’s functioning and collaboratively form a unified
overall process that is self-maintaining. This kind of holistic in-
terdependence of functioning of the parts makes the whole an
all-or-nothing unity (alive or dead)24.

In contrast to the all-or-nothing phenomenon of organiza-


tion, there can be degrees of cooperation. This is especially true
for hierarchically organized systems with subunits on various
levels of organization, for example, multicellular organisms. In
these systems there may be several levels on which systems of
interdependent parts can be identified, e.g., genes, cells, individu-

24 M.G. Miller, & A.E. Miller, Aristotle’s dynamic conception of the Psuchē as being-
alive, in S. FÖLLINGER (hg.), Was ist ,Leben’? Aristotelesʼ Anschauungen zur
Entstehung und Funktionsweise von Leben, Steiner, Stuttgart 2010, p. 69.
70 Georg Toepfer Articles

als, groups and ecological communities. In the field of biology it


has become practice to call at least some of the organized systems
on these levels ‘organisms’. Cells have been called «minimal
organisms» and groups of one species or ecological communities
«super organisms». But on the basis of the theory of natural
selection, it is also possible to locate the organism on only one
level, namely on the level where the systems are most integrated.
This is the way David Queller defines the organismic level, name-
ly as that level of integration in the biological hierarchy of entities
on which conflict (i.e., selection) between the parts is minimal
and cooperation maximal:

We designate something as an organism, not because it is n


steps up on the ladder of life, but because it is a consolidated
unit of design, the focal point where lines of adaptation con-
verge. It is where history has conspired to make between unit
25
selection efficacious and within-unit selection impotent .

On the organismic level powerful selection takes place be-


tween the units (e.g. individual organisms) and weak selection
within the units (e.g. the cells within one organism). This pattern
results in high levels of integration and cooperation in the units26.
As a consequence, in some cases Queller locates the organismic
level at the level of what has elsewhere been called the super-
organism of a society. For example, among eusocial insects (i.e., in
insect societies with sterile worker classes) the bee colony appears
on this level. Here cooperation is most significant with little
conflict between the bees.
The defining feature of the organism is the functional inte-
gration of its parts into one common goal: the maintenance and
reproduction of the whole system. At higher levels than the
organismic level there are, speaking in functional terms, cross-
purposes that are not integrated into one common goal.
In the major transitions of evolution, the organismic level has

25 D.C. QUELLER, Cooperators since life began, «Quarterly Review of Biology», 72,
1997, pp. 184-188, p. 187.
26 D.C. QUELLER, Relatedness and the fraternal major transitions, «Philosophical

Transactions of the Royal Society», B 355, 2000, pp. 1647-1655, p. 1653.


Articles Organismicity and Deconstruction of the Organism 71

moved up towards more complex systems. In these transitions,


there was not only an «evolutionary transition in individuality»
taking place, as Richard Michod has called it27, but also an «evolu-
tionary transition in organismicity» from lower to higher levels.
The key point of this discussion is that ‘organism’ is a rela-
tive concept. It designates systems that have undergone massive
changes in evolution and that were created by natural selection on
various levels. Organismicity is a product of evolution and comes
in degrees.

4. Agency in Degrees

Agency is a difficult term to define. Some recent attempts at


a definition include the following:
By autonomous agent I mean any embodied system designed
to satisfy internal or external goals by its own actions while in
continuous long-term interaction with the environment in which
it is situated28.
Agent systems are systems that can initiate, sustain, and
maintain an ongoing and continuous interaction with their envi-
ronment as an essential part of their normal functioning29. Agen-
cy occurs as soon as a system is able to modify some of its
boundary conditions ‘for its own sake’30. An agent is an autono-
mous organization capable of adaptively regulating its coupling
with the environment according to the norms established by its

27 R.E. MICHOD, Darwinian Dynamics. Evolutionary Transitions in Fitness and


Individuality, Princeton University Press, Princeton 1999; R.E. MICHOD & A.M.
NEDELCU, On the reorganization of fitness during evolutionary transitions in individuality,
«Integrative and Comparative Biology», 43, 2003, pp. 64-73.
28 R.D. BEER, A dynamical systems perspective on agent-environment interaction,

«Artificial Intelligence», 72, 1995, pp. 173-215, p. 173.


29 T. SMITHERS, Are autonomous agents information processing systems? in L. STEELS

& R.A. BROOKS (eds.), The Artificial Life Route to Artificial Intelligence. Building
Situated Embodied Agents, Erlbaum, Hillsdale 1995, p. 97.
30 K. RUIZ-MIRAZO, & A. MORENO, Searching for the roots of autonomy: The natural

and artificial paradigms revisited, «Comunication and Cognition-Artificial


Intelligence», 17, 2000, pp. 209-228, p. 217.
72 Georg Toepfer Articles

own viability conditions31.


In most definitions ‘agency’ refers to the regulation of a sys-
tem, especially the regulation of the coupling of the system to its
environment. In ordinary biological terms, this is called behavior.
Agency or behavior does not refer to the constitution of the system,
(i.e., the coupling of the system components) but to its maintenance
by regulating the coupling of the system to its environment. This
regulation presupposes, of course, the existence of a system to be
regulated. In this respect, ‘agency’ is a less fundamental concept
than ‘closure’ or ‘cooperation’. The term may also be applied to
sub-organismic levels such as physiological processes, neurody-
namic patterns or even chemical reactions. Thus, agency does not
have to be subordinated to biological organization. Indeed, many
authors admit that according to their definition there are also
agents in the inorganic world, e.g., synthetically produced «mini-
mal forms of proto-cellular systems»32. Therefore, agency does
not seem to be a very good candidate for a criterion in an unam-
biguous definition of the concept ‘organism’.

5. The Deconstruction of the Organism

With these various continua regarding central aspects of the


organism in mind – the continua in closure, cooperation and
agency – ‘organismicity’ instead of ‘organism’ seems to be an
attractive term. It adequately expresses the gradual nature of
various aspects of the organisms’ way of being, in particular the
following six:
(1) The transition from organized systems in nature to organ-
isms: There are inorganic systems in nature with interde-
pendent processes, like the water-cycle, or even pieces of
matter with mutually dependent parts that can reproduce
themselves, as is the case with clay particles. In these sys-

31 X. BARANDIARAN, E. DI PAOLO & M. ROHDE, Defining agency. Individuality,


normativity asymmetry and spatio-temporality in action, «Adaptive Behavior», 17, 2009,
pp. 367-386, p. 376.
32 Ivi, 367.
Articles Organismicity and Deconstruction of the Organism 73

tems, the mutual dependence of process types (functional


closure) can be found, while other central aspects of living
organisms, like the structural closure of the system in one
coherent body, are not present.
(2) The beginning and ending of an organism in time: Organ-
isms gradually come into being and gradually disappear: «Or-
ganismicity is achieved in steps and lost in steps at death»33.
It is unclear at which point in its development a human be-
ing starts being an organism: Is it with the coming into being
of a unique genetic code at fertilization of the egg cell? Is it
with the inability to divide into genetically identical twins at
the stage of the gastrula? Or is it when the fetus loses physio-
logical contact to its mother at birth?
(3) The spatial limits of individual organisms in cases of clonal
growth: Organisms with clonal growth like the American
trembling aspen Populus tremuloides or the fungus Armillaria
ostoyae can extend for hundreds of hectares. They are consid-
ered the largest organisms on earth; but it also seems reason-
able to see each tree (with its roots) or fruiting body (with
its mycelium) in these populations as an organism. In these
cases, like in the famous case of slime molds, it seems ap-
propriate to accept that there can be several levels of organ-
ismicity within one species. Important life-functions such as
metabolism or reproduction are realized at the level of the
parts, the individual trees or fruiting bodies, but the system
operates more effectively when it keeps the parts of the en-
tire complex together.
(4) The focal level of organization within an organism: In
multicellular organisms there are levels of intraspecific or-
ganismicity that may extend from cells and (morphologi-
cal) individuals to familial groups of super-organisms and
populations.

33J.J. DIAMOND, Abortion, animation and biological hominization, in Abortion, part


IV, Hearings before the Subcommittee on Constitutional Amendments of the
Committee on the Judiciary, Ninety-forth Congress, First Session, U.S. Gov.
Print, Washington D.C 1976, pp. 594-605, p. 602.
74 Georg Toepfer Articles

(5) The distinction between organism and environment: In


some organisms, such as spiders and their webs or corals
and their substrate, it is not clear whether external entities
produced from bodily substances or external chemical re-
actions catalyzed by the metabolism of the organism be-
long to the organism or its environment. On the one
hand, they are produced by the organism and feedback on
its functional needs, but on the other hand they are not
part of the morphological unity of the system. As organ-
isms tend to construct their ecological niches and to inte-
grate parts of their environment into their functional sys-
tem, their organismicity can gradually extend beyond their
morphological boundaries 34.
(6) The transition from individual to community: Even for
entities within an organism it is not always clear whether
they are part of the organism or not. One example would
be the bacterial symbionts in our body: they depend on us,
and we depend on them. This means that there are two
types of organisms: one intraspecific organism of Homo sa-
piens and another transspecific organism, which is Homo
sapiens together with its symbiotic community of bacteria.
Which of these is considered the more integrated system
depends on the context.
For each of these facets, it seems to be more appropriate
to speak of ‘organismicity’ or ‘organismality’ instead of ‘organ-
ism’. A shift in the term ‘organism’ to this gradual sense would
mean a return to the pre-Kantian definition of the word, which
referred to the causal structure of a system rather than the
system itself. But, even if the gradualist term ‘organismicity’
seems to be fitting in many contexts, it is still possible to main-
tain that organisms do indeed exist in nature and to accept
them as ontologically real, especially in prototypical cases like
human beings. They are systems of interdependent parts that
have a body produced by the system itself and which in turn

J.S. TURNER, The Extended Organism, Harvard University Press, Cambridge,


34

Mass. 2000.
Articles Organismicity and Deconstruction of the Organism 75

has a constraining influence on the processes maintaining the


system and its reproduction. The concept of organismicity
simply stresses the point that all these characteristics, except
the relation of interdependence, admit of degrees. Given that
organisms are systems shaped by the gradual process of natural
selection, it would be surprising to observe this to be otherwise.
THE RISE AND FALL OF THE MACHINE METAPHOR:
ORGANIZATIONAL SIMILARITIES AND DIFFERENCES
BETWEEN MACHINES AND LIVING BEINGS

by Victor Marques & Carlos Brito

Abstract. Our goal in the paper is to offer both an eulogy and a critique of
the machine metaphor as a theoretical resource for understanding organic
systems. We begin by presenting an abbreviated history of the machine meta-
phor, pointing out how it was instrumental in the development of modern
biology, as it provided a conceptual basis for an analytical program in the
sciences of life. Then we deal with what exactly makes the machine metaphor
such a successful resource, pointing to what organisms and machines in fact
share in common – based on the relational approaches advanced by Varela
and Rosen, we suggest that both are ʻconstrained systemsʼ. In the third part,
we present an alternative way of conceptualizing living systems, bringing now
the disanalogies with machines to the foreground. Reviewing the independent
work of different authors, we show that there is distinct organicist theoretical
camp, where the organism is generally understood as an autonomous system.
Finally, we observe that many authors from that camp are now reclaiming
Kant’s treatment of organisms in the Critique of Judgment, in particular the
concept of «natural purpose» – but those authors do that with a markedly
anti-Kantian goal: to naturalize teleology. Our conclusion is that the view of
organism as an autonomous system gives us the key to a naturalistic under-
standing that can finally overcome the mechanical view of nature so character-
istic of modern thought. The machine metaphor, despite all its undeniable
contributions to the advancement of biological research, shows itself ultimately
insufficient for a complex view of the phenomena of life – and discarding it
doesn’t need to mean any concession to vitalism: on the contrary, it may be
exactly what we need to invigorate a robustly materialist project.

1. The Machine Metaphor

It is possible to trace the origin of the machine metaphor


back to the Middle Ages, particularly to the time when technolog-
ical advances and the production of artifacts of increasingly
Verifiche XLIII (1-4), 2014, pp. 77-111.
78 Victor Marques & Carlos Brito Articles

complexity exerted profound cultural influence and provided new


images to interpret the natural world. Mechanical parlance served
well the pioneers of the systematic study of anatomy and physiol-
ogy to understand the new data in terms of already familiar expe-
riences. Technology therefore offered an intuitive picture of
nature upon which science could work. With the emergence of
the mechanical philosophies of the seventeenth century, and the
ambition to give an account of the whole of nature in terms of
inert matter interactions alone, it was only natural to think of life
as nothing more than a specific type of machine, the difference
between organisms and mere artificial automata reduced to a
quantitative one, residing solely in the degree of complication.
One should thus place the effort of developing a mechanical
paradigm in the context of the historical emergence of modern
science, with the successful appearance, especially after Galileo,
of a new physics, opposed to classical Aristotelian physics and in
an ongoing struggle with the animistic worldview.
One of the most conspicuous aspects of the conception of
nature presented by Aristotle is the centrality the notion of final
cause plays within it. Its teleological character resides precisely in
the claim that natural objects have a function, expressing thus
purpose. For Aristotle, that purpose is immanent in reality itself
was clearly demonstrated, for example, in the natural tendency of
living beings to develop, regenerate and respond adaptively to
environmental influences. Organic entities were considered there-
fore both material and teleological systems, at the very same time.
Aristotle was not only one of the greatest thinkers of antiquity
but also one of the first, if not the very first, biologist, and his
vision of Cosmos was based after the model of the organism1.
In fact, some type of ʻpan-vitalismʼ seems, as pointed by
Hans Jonas2, to be the primitive hypothesis of the human inter-
pretation of Being, pre-modern way of thinking being marked

1 As noted by Lenny Moss: «Nature as a whole for Aristotle was lifelike –


conceptually modeled not by the example of inertness but rather by the
example of living activity». L. MOSS, What Genes Can’t Do, MIT Press, Cam-
bridge 2004, p. 7.
2 H. JONAS, Das Prinzip Leben, Suhrkamp, Frankfurt 1977.
Articles Rise and Fall of the Machine Metaphor 79

thus by a kind of ontological dominance of life. Properly modern


thought, on the other hand, is distinguished precisely by the very
opposite movement:

This kind of outlook changed dramatically during the metaphys-


ical shift that took place over the course of the seventeenth cen-
tury. Nature became stripped of its capacity to self-organize as
an end unto itself3.

The physics of Galileo differs from the Aristotelian physics


most notably by the complete absence of final causality. The con-
ceptual shift can be viewed as follow: what Aristotle intended to
explain through a natural teleology may now be better understood
in terms of efficient causality and nothing more. Final causes
becomes unnecessary, and reliance on them, unscientific in princi-
ple. In losing its generic status, and demoted, from something
nearly universal, to a position of a particular case (a very special
one), life must then be explained in terms of non-living phenome-
na, that is to say, embedded in the theoretical space now monopo-
lized by the hypothesis of comprehensive pan-mechanicism.
The scientific program of mechanization of living phenome-
na, as a constitutive part of a naturalistic mechanical project,
actually starts to gain traction from the seventeenth century
onwards, and especially with the publication of Harvey’s treaty on
circulation. Here we already see the analogy with machines paying
off theoretical and practical outcomes, helping to produce prom-
ising research hypotheses and lines of inquiry. The actual concep-
tualization of a philosophically grounded and generalized ma-
chine metaphor is commonly attributed to Rene Descartes.
Robert Rosen, a theoretical biologist and key figure in the con-
temporary discussions on the machine metaphor, locates its roots
in the analogy with artificial toy automata common in the royal
gardens of 17th century:

The machine metaphor was first proposed by Descartes in the


early seventeenth century. It is reported that, as a young man,
Descartes was much impressed by some lifelike hydraulic au-

3 L. MOSS, What Genes Can’t Do, cit., p. 7.


80 Victor Marques & Carlos Brito Articles

tomata. With characteristic audacity, he later concluded from


these simulacra that life itself was machinelike 4.

The power and intellectual appeal of the metaphor stem


from the fact that complicated machinery in general serves as a
very practical demonstration that purely mechanical systems,
when properly organized, can perform all kind of complex behav-
iors. Even more: we know exactly how the machine works – after
all, we built it. We know precisely what it is because we con-
structed it; it is epistemically transparent. We can then be confi-
dent that there is nothing magical or mysterious in the operation
of a machine, no matter how amazing or counter-intuitive its
competences are – nothing at all that doesn’t follow strictly the
very universal and basic laws of physics that apply to inanimate
matter, being thus fully reducible to mere efficient causality.
Therefore, it is not necessary to postulate any vital principle or
inner hidden telos to explain how a machine does what it does.
And if by means of a machine we can mimic some characteris-
tics and behaviors of living beings, nothing prevents us to even-
tually build such a machine capable to emulate all of them – the
difference between one and the other would be only one of
degree, not of quality. Just as there is nothing miraculous in the
machine, there is also no need for something miraculous in the
organism either.
It is easy to see, however, how the machine metaphor lends
itself naturally to theological speculations. What is now miracu-
lous is not that the organism functions, but that the organism
exists. As the argument used to go: from a machine it can be
deduced the existence of a machine constructor, a designer. The
internal teleology inherent in the natural entity is discarded, but
just to be replaced by the external teleology derived from a ra-
tional mind, which organizes the material from the outside so that
it acquires functionality: «Final cause, the for-the-sake-of-which a
creature possessed the form that it comes to have, was not lost
but rather relocated. Seventeenth century metaphysicians moved

4 R. ROSEN, Essays on Life Itself, Columbia University Press, New York 2000, p. 266.
Articles Rise and Fall of the Machine Metaphor 81

final cause from within nature to the mind of God»5.


Incidentally, this compatibility with Christian theology has to
be counted as one reason for the success of the metaphor. At the
same time that it freed scientists from direct theological consider-
ations when investigating physical nature (since material systems
themselves, other than human beings, are no longer inherently
sacred), it also reserves an untouchable place for God and spirit –
outside nature, and thus also outside the scope of scientific inves-
tigation. Nature becomes despiritualized, disenchanted, but by the
very same movement, spirit becomes dematerialized, disembod-
ied – two absolutely distinct kinds of stuff. Kampis captures
precisely this ambiguity in the mechanical wordview:

Mechanistic materialism, a philosophical pedestal for mechanis-


tic thinking, is deistic. […] Deists assume God’s existence as crea-
tor but do not let him intervene in the world he has created. This
was reflected in nineteenth century thinking, admittedly mecha-
nistic but also religious, according to the spirit of the time6.

The mechanistic monism in the realm of the natural world,


while entering into confrontation with a pre-modern conception
of a spiritual, lively, spontaneously productive nature, reinforced
another religious view, that of a transcendent creator, who by his
own will externally imposes form on a exteriorized, indifferent
matter, in itself essentially inert and formless. Since the new
physics was based on a non-teleological conception of matter, all
finality had to be externalized, expelled from nature, thereby
replacing the immanent teleology by a transcendent teleology.
The final, fatal blow to teleology would have to wait for Darwin
to be finally delivered. The final form of the “Enlightned” theory
of organism, dominant during most of the 20th century, was a
combination of the machine metaphor plus a Darwinian view of
evolution, with natural selection acting as a kind of “blind
watchmaker” designing “moist robots”.
The machine metaphor brought us undeniable intellectual

5 L. MOSS, What Genes Can’t Do, cit., p. 7.


6 G. KAMPIS, Self-modifying systems in biology and cognitive science. A New Framework
for Dynamics, Information and Complexity, Pergamon Press, Orford 1991, p. 192.
82 Victor Marques & Carlos Brito Articles

benefits. It not only brings biology closer to engineering (and to


physics, as a consequence), but it also offers a thoroughly natural-
istic model, a general framework for explaining the activity of
living beings without recourse to any principle beyond the scope
of science research and potential empirical investigation. To a
large extent, the metaphor owes its vitality and popularity to the
privileged role in the battle against vitalism. For the concrete
practice of the scientist, however, it has had an even more im-
portant methodological consequence, as it grounds, and serves as
the basis for, an analytical research program in biology. The
machine metaphor accomplishes that by providing the theoretical
justification for assuming that the same procedure used success-
fully to study machines in general is also applicable to study any
organic system: dismember it into its constituent parts and char-
acterize them individually as independent sub-systems7. To the
extent that the organisms are similar to machines, the analytical
strategy works: biological research can thus accumulate
knowledge and progress forward by treating living beings as a
complicated ensemble of discrete mechanisms, describing par-
ticular components and identifying particular functions.
There is no denying that this program has been immensely
successful. The analytical strategy based on the machine meta-
phor remained at the center of the spectacular advances in, for
example, molecular biology during the last century. There would
be no way to explain such astounding success if there were not in
fact real similarities between machines and organisms. If practice
is the proof of theory, it’s mandatory to accept that, at least in
some sense, organisms can be viewed as machines. But in what
sense? Or, what exactly makes the machine metaphor such a
good metaphor?

7 As stressed by Kampis: «[…] machines are decomposable in the sense that


they are built from stable and separately accessible parts, which have separately
knowable properties. Much as the machine is made up from these separate
parts, so is the understanding of the machine made up from pieces of under-
standing of its atomistic parts». G. KAMPIS, Self-modifying systems in biology and
cognitive science, cit., p. 186.
Articles Rise and Fall of the Machine Metaphor 83

2. Organisms and machines as constrained material systems

We now know very well that living beings are composed of


the same kinds of atoms that are found in other parts of nature,
and no other special or unique substance enter in their making –
organisms are made out of the exact same stuff as the rest of the
entire, life-less, universe. But unlike most of the physical or chemi-
cal systems which we encounter in the inanimate world, living
beings are very complex, consisting of great number of parts,
diverse in form and shape, arranged in a specific and intricate
fashion: to sum, organisms are material systems both heterogene-
ous and ordered. Additionally, each part seems to express a definite
purpose, as if it was a product of design. Organisms, of course, are
not the only material systems to express all those characteristics:
also in the case of machines each one of them is present.
In fact, as physical entities, both organisms and machines are
systems marked by certain internal heterogeneity, amidst which is
possible to distinguish differentiated parts. Also, both machines
and organisms are organized into functional component, thereby
allowing a description of the overall behavior making use of no-
tions such as functioning, regulation, operation and coordination.
On those grounds, Chilean biologists Humberto Maturana
and Francisco Varela, following the anti-vitalist tradition, identi-
fied organisms as machines8. By emphasizing the mechanic nature
of living beings they wanted above all to assert two basic com-
mitments. First, their firm belonging to modern scientific tradi-
tion, explicitly rejecting any residue of animism, denying that
there is anything magical or incomprehensible in the passage
from non-life to life, no forces or principles whatsoever which
are not found in the material universe – living beings in no way
escape from the ordinary physical laws that are valid to inanimate
systems. Equally important, however, is that, as Varela specially is
keen to stress, just as machines, organisms are defined by their

8 H. MATURANA & F. VARELA, Autopoiesis and Cognition. The realization of the


living, D. Reidel Publishing Company, Dordrecht 1980.
84 Victor Marques & Carlos Brito Articles

organization9. In this sense, what is really characteristic of ma-


chines is that they are coordinated systems of components capa-
ble of satisfying certain relations.
By adopting such abstract and most general account of ma-
chine, Varela is making a functionalistic point. He wants to draw
attention to the relative independence of an abstract organization
with respect to the substrate, in sum, to a principle of multiple
realizability: a machine can be materialized through various physi-
cal structures, the precise, detailed material nature of components
being largely irrelevant given that they are able to play specific
functional roles – in short, it doesn’t matter the stuff the compo-
nent is made of, only what it is capable of doing (under certain
conditions). That is to say: what is really relevant about a compo-
nent are the relational dispositions in interacting with other com-
ponents. The organization of a machine, defined by its relations,
is logically independent of the properties of the components,
which are, to a certain extent, arbitrary: «a given machine can be
realized in many different ways by many different kinds of com-
ponents»10. Therein lies the distinction drawn by Varela between
organization and structure – organization being a purely relational
notion, with no connection to materiality, and structure the
concrete physical realization.
In fact, that very same distinction had already been articulat-
ed and put into play by Robert Rosen in his own relational treat-
ment of life11. Rosen contrasts models of a «structural character»
to «higher level» models, which deal with the «functional organi-
zation», and then calls attention to the same principle of multiple
realizability, stressing that «essential features of cellular organiza-

9 «In saying that living systems are ʻmachinesʼ we are pointing to several
notions that should be made explicit. First, we imply a nonanimistic view,
which should be unnecessary to discuss any further. Second, we are emphasiz-
ing that a living system is defined by its organization, and hence that it can be
explained as any organization is explained, that is, in terms of relations, not of
component, properties». F. VARELA, Principles of biological autonomy, Else-
vier North Holland, New York 1979 , p. 7.
10 Ivi, p. 9.
11 R. ROSEN, Some relational cell models: the metabolism-repair systems, in Foundations

of Mathematical Biology, Academic Press, New York 1973.


Articles Rise and Fall of the Machine Metaphor 85

tion can be manifested by profusion of systems of quite different


structure». In order to complement a merely «structural study of a
biological system», which abstracts away the organizational prop-
erties, he calls for the necessity of the opposite strategy, abstract-
ing away the structure and leaving behind only the organization,
to be studied abstractly in purely theoretical terms, divorced from
any particular realization. That will become his later duality be-
tween a reductionist approach – throw away the organization and
keep the underlying matter – and a relational approach – throw
away the matter and keep the underlying organization12.
It’s now easy to see that both Rosen and Varela were work-
ing in the context of the developing movement of theoretical
biology (inspired by ideas from cybernetics and systems theory)
and the attempts to formulate a general theory of organized sys-
tems, understood as «the study of organization per se, divorced
from material embodiment»13. A relational approach allows thus
one to treat organization as «a thing in the abstract», making visible
the arbitrary relation between an organization and the specific
materiality of its components: «there is nothing in the components
that mandates that particular organization, nor anything in the
organization that mandates those particular components»14. Both
Rosen and Varela were after what is special about life not in terms
of substance – of what it is made of – but in terms of pure form –
of how it is organized. Which is precisely why Varela in his Princi-
ples of biological autonomy uses ‘machines’ and ‘systems’ interchangea-
bly: «machines and systems point to the characterization of a class
of unities in terms of their organization»15. That also explains why
both Rosen and Varela, at least in their earlier works, claim to be
advancing mechanistic theories of life16.

12 R. ROSEN, Life Itself. A Comprehensive Inquiry into Nature, Origin and Fabrication
of Life, Columbia University Press, New York 1991.
13 Ivi, p. 14.
14 Ivi, p. 140.
15 F. VARELA, Principles of biological autonomy , cit., p. 7.
16 «Relational biology itself is as mechanistic a theory as any reductionist

structural theory»
R. ROSEN, Some relational cell models: the metabolism-repair systems, in Foundations of
Mathematical Biology, New York 1973.
86 Victor Marques & Carlos Brito Articles

It is not in the least surprising thus that Rosen points to the


very same similarity between machines and living beings:

What distinguishes a material system as a machine, as distinct


from a stone or a crystal, must somehow reflect its intrinsic or-
ganization. [...] But once we talk about organization, we are in a
relational context. We are basically defining machine as a material
system that admits (i. e., that realizes) a relational description17.

For Rosen, a system may be called “organized” to the extent


that it is possible and useful to analyze it as consisting of compo-
nents, which in turn are defined in functional terms from their
role in the relation to the other components of the system. The
component is the organizational unit: a part with a function. The
relationships between the components in determining the or-
dered activity of the system as a whole is the object of what
Rosen calls «relational theory of systems». What machines and
living beings share in common is that they are both material
systems that admit relational descriptions.
What we are dealing with here is nothing like the “organizing
force” of vitalism, a “something more” added to mere materiality,
but, as Ashby highlights, with organization as a coordinated
system of constraints:

in the past, biologists have tended to think of organization as


something extra, something added to the elementary variables,
the modern theory, based on the logic of communication, re-
gards organization as a restriction or constraint18.

Organisms, however, just as man-made machines, are not


pure organization, but actual material systems realized through

«By adopting this philosophy, we are in fact just adopting the basic philosophy
that animates cybernectics and systems theory […]. This is, I believe, nothing
more and nothing less than the essence of a modern mechanicism». F. VARELA,
Principles of biological autonomy, p. 7.
17 R. ROSEN, Life Itself, cit., p. 183.
18 W.R. ASHBY, Principles of the self-organizing system, in Principles of Self-

Organization: Transactions of the University of Illinois Symposium, ed. H. Von Foer-


ster and G. W. Zopf, Jr. Pergamon Press, London 1962.
Articles Rise and Fall of the Machine Metaphor 87

physical structures. So how then to pass from an abstract charac-


terization to concrete existence, or, how to think the physical
realization of constraints? In his classic paper from 1968, Life’s
irreducible structure, Polanyi seems to point to the right direction19.
Polanyi observes that the constructor of a machine «restricts
nature in order to harness its working». But how does this re-
striction take place, and becomes physically effective? By impos-
ing boundary conditions on the laws of physics and chemistry.
And the same is also valid for living beings, whose components
perform functions in the same way as parts of a machine do:

in this light the organism is shown to be, like a machine, a sys-


tem which works according to two different principles: its struc-
ture serves as a boundary condition harnessing the physical-
chemical processes by which its organs perform their functions.

It is useful to note the distinct character of these two principles:


while the laws of nature are, in principle, universal, inexorable,
and incorporeal, the constraints are, by definition, local and
arbitrary and require a specific physical realization, they are also a
product of physical processes and exist in a determinate context –
they are not only constructed by a previous process but precari-
ous, vulnerable to destruction and natural decay. Constraints are
physically realized by way of boundary conditions harnessing the
laws of inanimate nature in order to do something (work, e.g.),
and both machine and organisms are better characterized as
material systems of coordinated constraints. They are constituted
by components that canalize physico-chemical processes in order
to realize functions. As Rosen puts it, machines and organisms
are «constrained systems»20.
What associates organisms and machines is that in both cases
we are dealing with materials systems in which the organizational
aspect is crucial and essential – when we want to understand the
machine as machine, the relevant thing to offer is a relational
description (a purely material description, albeit still possible, will

19M. POLANYI, Life’s irreducible structure. «Science», LIV (160), 1968, p. 1308.
20R. ROSEN, Causal Structures in Brains and Machines, «Int. J. General Systems»,
LIV (12), 1986, p. 107.
88 Victor Marques & Carlos Brito Articles

inevitably overlook what is actually relevant about the phenomena


in question). And the same goes in the case of organisms. Without
that central underlying communality, the machine metaphor would
never work. As far as it is theoretical fruitful, it’s because we are
dealing fundamentally with organized entities.
That is why the machine metaphor has proved so powerful
and long lasting, to the point that it has managed to survive even
the crumbling of two other notions that used to accompany it,
and forming together a coherent theoretical system: the idea of
special creation and preformationism. The recognition of the
historical character of living beings and the understanding of
development as a process of epigenetic nature seem to weaken
the metaphor, forcing important disanalogies, since, as a rule, the
appearance of a machine is connected to the activity of a con-
scious mind and does not involve the spontaneous emergence of
order as in development, the process of differentiation that does
take place during biological ontogeny. In fact, if organic bodies
are machines, it is clear that they are a very peculiar kind of ma-
chines, which human beings are not yet able to build.
Howard Pattee (1971) was one of the first to conceptualize
organisms as coordinated sets of constraints, and used this notion
to give precise expression to the idea of interactions among
different hierarchical levels of organization21. More recently,
authors such as Kauffman (2000)22, Mossio and Moreno (2010)23,
and Deacon (2012)24 are reclaiming, in a way or another, the
concept of constraint in order to make explicit the specificity of
living beings. It is only at this level of abstraction that the essen-
tial difference between organisms and machines shows itself.
Meanwhile the components and organization of a machine are
defined from the outside, with its constraints externally imposed
upon the material structure, in the organism the opposite is true,

21 H.H. PATTEE, Physical theories of biological co-ordination, «Quarterly Reviews of


Biophysics», LIV (4, 2 & 3), 1971, p. 255.
22 S.A. KAUFFMAN, Investigations, Oxford University Press, Oxford 2001.
23
M. M OSSIO & A. M ORENO , Organisational Closure in Biological Organisms.
«Hist. Phil. Life Sci.», LIV (32), 2010, p. 269.
24 T. D EACON , Incomplete nature: How mind emerged from matter , Norton

& Company, New York 2012.


Articles Rise and Fall of the Machine Metaphor 89

since its constraints are permanently posited, generated, repaired


and modified as a result of the system’s own functioning. Organ-
isms and machines are constrained systems, but living beings are,
on the top of that, also autonomous systems that produce their
own constraints; that is, the constrained dynamics of the lower
level elements results in the maintenance and replacement of the
very constraints that control their behavior, in a closed loop.
Terrence Deacon, also recognizing that both machines and
organisms are functionally organized systems, calls attention to
important differences between them. While machine parts are
produced separately and only put together later during the as-
sembling process, nothing similar happens in the organic case: the
organism develops spontaneously, its parts differentiating from
relatively undifferentiated starting point, with no exterior assem-
bling needed, the components grow out of a prior unity, coming
to be already integrated with the ever functioning whole. Organic
components are from the beginning interdependent and involved
in multiple relations with other components, both contributing to
the function of the rest of the system and being reproduced by
the networks of organic processes. On the other hand, with
conventional machines, the manufacturing process is distinct
from the operation process: the machine is constructed out of
pre-existing components, with purpose of performing a function
that is generally not related to its own production25.

25 Nicholson (2013) points roughly to the same direction. He acknowledges


some elements of similarity between organisms and machines – both are
physical systems that function in accordance to natural laws, consume energy
and transform part of it into work, are hierarchically structured and internally
differentiated, admit relational description and can thus be modeled in terms
of causal relations between interacting components, and are fruitfully charac-
terized in teleological terms – but stresses that all those similarities are ulti-
mately superficial, and the distinctive features of organism are fundamentally
different from those of machines. Above all, «organisms are intrinsically
purposive, whereas machines are extrinsically purposive». Nicholson notes that
in a machine the parts are causally independent, and temporally antecedent, to
the whole, whereas in an organism neither is the case: the parts exist in a
relation of collective interdependence – «the organism maintains its autonomy
as a whole by constantly regulating, repairing, and regenerating its parts». He
also draws attention to the peculiar «transitional structural identity» displayed
90 Victor Marques & Carlos Brito Articles

In the organism we are faced with a material system in which


the components depend for their very existence on the overall
context of organic activity, within which they are continuously
being reproduced and repaired. Each functional component is
materially dependent on the context afforded by the joint and
coordinated activity of the other components, in a meshwork of
tangled mutual dependence and reciprocal relations. In fact, we
can say that it is the organism itself that defines its own limits,
that is to say, the constraints that give it a definite shape are
posited in and by the actual process of functioning of the organ-
ism as a articulated whole, the result of its activity being thus its
continued existence.
In addition, the organism constituents are liable, plastic, in-
cessantly changing, constantly being modulated by the metabo-
lism to ensure adaptive responses. The regularity and stability of
the organism does not result from materials that resist defor-
mation, its permanence is not rooted on material inertness. Un-
like a machine, that maintains its form because the physical con-
stitution of its components is such as to render the interactions
between them relatively insensitive to thermodynamical fluctua-
tions, organisms exploit non-equilibrium chemical processes
generating order by self-organization.
Just as the machine metaphor has several advantages that
helped foster scientific progress, it also has some important
limits, that, if not properly acknowledged and explicitly criticized,
can severely impoverish the conceptual space in which we make
sense of complex biological phenomena. As Lewontin & Levins
expressed it,

the machine metaphor creates a general program for biological


investigation that is circumscribed by just those properties that
organisms have in common with machines, objects that have ar-

by the organism – the materials change, yet the organization remains: «While a
machine always consists of the same material components […] an organism
naturally maintains itself in a state of continuous flux in which there is a
permanent breaking down and replacement of its constituent materials». D.J.
NICHOLSON, Organisms ≠ Machines, «Studies in History and Philosophy of
Biological and Biomedical Sciences», LIV (44-4), 2013, p. 669.
Articles Rise and Fall of the Machine Metaphor 91

ticulated parts whose motions are designed to carry out particular


functions. So the program of mechanistic biology has been to de-
scribe the bits and pieces of the machine, to show how the pieces
fit together and move to make the machine as a whole work, and
to discern the tasks for which the machine is designed26.

3. Life as autonomy

In fact, if organisms are machines in this general sense, it is


clear that we are then dealing with a very unusual kind of ma-
chine. The common characterization of living beings as “reproduc-
ing machines” is ambiguous, as it can give rise to two, quite distinct,
forms of understanding the phenomena of life. On the one hand,
“reproduction” could be understood as replication, that is, pro-
duction of another entity with similar characteristics: the organ-
ism reproduces in the sense of multiplying itself, increasing in
numbers in a given population (that is the sense that first comes
to mind, for example, when we think about “selfish genes” repli-
cating – that is to say, making copies of themselves). The most
important fact about biological systems would therefore be that
they are the result of a reproductive event and therefore have a
genealogical history, which can also be extended into the future,
as a tree of successive reproductive events expanding indefinitely.
A point of key importance here is that from the notions of repli-
cation, inheritance, variation, and differential viability it is possi-
ble to deduce the principle of evolution by natural selection.
According to this view, what essentially characterizes life is its
connection with the evolutionary process: organisms are that
which, at the population level, can undergo a process of evolution
by natural selection and, at the individual level, are the result of a
prior evolutionary process. Evolution by natural selection is thus
viewed as the general notion that unifies biology, the universal
aspect of life and the basis where to ground the distinction be-
tween the mere physical world and the biological realm. Here the
idea is that the cumulative effect of natural selection over genera-

26R. LEWONTIN & R. LEVINS, Biology under the influence. Dialectical essays on ecology,
agriculture, and health, Monthly Review Press, New York 2007, p. 222.
92 Victor Marques & Carlos Brito Articles

tions results in a kind of natural design, that can be viewed in fact


as analogous to rational, intelligent design, conferring functions
to the parts of the organisms, and giving thus the appearance that
those parts were made for certain purposes.
However, the term «reproduction» can also be related to self-
production, or the active conservation of a self-established identi-
ty. A reproductive system is thus a system capable of self-
maintenance. This second sense of reproduction focuses not on
the genealogical history of the organism, or its potential for
further self-replication, but on the metabolic processes that at
every moment regenerate and/or modify the form of the living
being and its way of being in the world. Humberto Maturana and
Francisco Varela were pioneers in this second approach. They
called living organisms «autopoietic machines», that is, a machine
organized as a system of processes of production of components
concatenated in such a way to continually reproduce the topology
of the whole network of processes, and through its own activity
delimiting itself physically as a unit in space.
According to Maturana and Varela a crucial distinction can
be traced between autopoietic machines, on one hand, and allo-
poetic machines, on the other:

[...] in a man-made machine in the physical space, say a car,


there is an organization given in terms of a concatenation of
processes, yet, these processes are not processes of production
of the components which specify the car as a unit since the
components of a car are produced by other processes, which are
independent of the organization of the car and its operation27.

Allopoetic machines produce by their activity only something


different from themselves, while in autopoietic machines there is
a coincidence between operation and fabrication – the autopoiet-
ic machine fabricates itself as result of its own functioning.
Maturana and Varela pointed out that as a consequence of their
autopoietic way of being, a living system is autonomous, that is to
say, characterized by its own dynamics, which the environment

27 H. MATURANA & F. VARELA, Autopoiesis and Cognition. The realization of the


living, cit., p. 79.
Articles Rise and Fall of the Machine Metaphor 93

can «irritate», affect positively or negatively, but not determine. In


every interaction with what is outside, the organism responds in
its own way, subordinating the changes in its own structure to the
conservation of its autopoietic organization.
By «autopoietic organization» it is meant simply that the con-
stitution of the system is the result of a specific way to concatenate
internal processes, in which the overall result is the maintenance of
the conditions of existence for the system as a whole, as a unity.
This circular concatenation of processes allows the emergence of
an individuality, a processual, precarious identity that persists
through time despite a never ending flux of material composition
and all kinds of deformations caused by its necessary interaction
with the environment. The living system is a concrete unity, dy-
namically stable, maintained by a constant material and energetic
flow, and has a history of interaction with the environment which
both enables and threatens its existence as a distinct entity.
About the same time the theory of autopoiesis was being
formulated, the mathematical biologist Robert Rosen expressed
some ideas very similar to those developed by Maturana and
Varela. Rosen also highlights the difference between organization
and structure, and denies the hypothesis that what is essential to
life can be recovered by the accumulation of purely structural
investigations (what he calls «the hypothesis of reductionism»)28.
In particular, Rosen observes that the lifespan of a cell exceeds
considerably that of their components, and that there is, for each
functional component, a natural tendency to decay, as the physi-
cal structures spontaneously degrade to the point that they are no
longer able to perform their proper function within the system.
The system as whole, however, keeps functioning by continuous-
ly repairing and reproducing its components. The question of
«who repairs the repairers» promptly announces the threat of an
infinite regress. Rosen’s solution is to bend the chain of compo-
nents onto itself, including the reparative activity in metabolism
itself; to break that infinite regress, Rosen folds back the linear
hierarchy into a loop. In Rosenean parlance, the system becomes

28 R. ROSEN, Some relational cell models: the metabolism-repair systems, in Foundations


of Mathematical Biology, ed. Robert Rosen, Academic Press, New York 1973.
94 Victor Marques & Carlos Brito Articles

now «closed to efficient causation», expressing thus a distinct


circular character. The system as a whole expresses a self-
repairing property due to its circular organization. As noted by
Letelier et al (2003):

Rosen’s main result is the demonstration that the synergy of


metabolic and repair actions can imply, under some circum-
stances, self-replication in the sense of self-production (or self-
maintenance) of the complete metabolic network 29.

In a way similar to autopoiesis, in Rosen’s MR-systems «the


closed-loop hierarchy produces an integrated whole that has an
identity that can be perturbed by, but remains distinct from, its
ambience»30.
It is always possible to analyze separately the operation of a
particular segment of metabolism just assuming the operation of
the parts concerned as something given, but in the context of the
organism, and its whole metabolic network, it is necessary to
remember that all these parts are also always a result of metabo-
lism – ultimately, each component is produced by and through
the combined activity of other components. It is to highlight that
fact that Rosen makes a point of stating that organisms are cate-
gorically distinct from machines because they are «closed to the
efficient cause», a property that can only be expressed by impre-
dicative models. Organisms are closed in the sense that every
efficient cause, which for Rosen means the function that controls
a material transformation, is itself in turn produced within the
organism. The contrast with manmade conventional machines is
obvious, since in these, generally, each component is separately
produced and subsequently put together in an ordered, but exter-
nal, fashion. Not only the mechanisms that produce them are
exterior to the artifact being produced, but also the very opera-
tion of the artifact has, in general, nothing to do with the re-

29 J. LETELIER, M.L. CARDENAS, A. CORNISH-BOWDEN, From L’Homme Machine


to metabolic closure: Steps towards understanding life, «Journal of Theoretical Biology
», LIV (286), 2011, p. 100.
30 S. W. KERCEL, The Endogenous Brain. «Journal of Theoretical Biology», LIV

(3-1), 2004, p. 61.


Articles Rise and Fall of the Machine Metaphor 95

pair/maintenance or replacement of components – the artifact


activity has nothing to do with its production process31.
Varela’s notion of organizational closure is thus related to
Rosen’s notion of a system being «closed to efficient causation».
Although the two models have been developed independently,
both seem to have a similar goal: to emphasize the importance of
circularity in modeling the phenomena of life and to affirm a
concept of organism based on circular organization. Both Rosen
and Varela are mainly interested in the basic question of a general
theory of biology: what makes living things alive? What do all
organisms share that separate them from inanimate matter? Or,
as Rosen poses the question: what is life itself? Investigating
living systems from a relational perspective, both Varela and
Rosen come to the conclusion that the answer to the question of
what is life itself lies in its circular organization: organisms are
natural realizations of «strange loops»32.
Hofstadter coined the term «strange loop» to make reference
to situations in which in a series of stages that constitute a cy-

31 Following Rosen ‘s analysis, Hofmeyr concludes: «the defining difference


between a living organism and any non-living object should be that an organism
is a system of material components that are organized in such a way that the
system can autonomously and continuously fabricate itself». J-H. S. HOFMEYR,
The biochemical factory that autonomously fabricates itself: A systems-biological view of the
living cell, in Systems Biology: Philosophical Foundations, ed. F.C. Boogerd, F. Brug-
geman, J.-H.S. Hofmeyr and N.V. Westerhoff. Elsevier, Amsterdam 2007, p. 217.
32 «Strange loop» is a concept coined by the cognitive scientist Douglas Hof-

stader, but Varela, already in 1984, also used it in an article to express that with
the phenomena of autonomy we enter in a “world of strange loops”: «A loop
is completed whereby two levels are collapsed, intercrossed, entangled. At this
point, what we wanted to hold in separate levels is revealed as inseparable, our
sense of direction and foundation seems to falter, and a sense of paradox sets
in». F. VARELA, The Creative Circle: Sketches on the Natural History of Circularity. in
The Invented Reality. How do we know what we believe we know? (Contributions to
constructivism), ed. P. Watzlawick. W.W. Norton & Company, New York 1984, p.
309. Deacon also mentions it, noting that biomolecules exhibit «process-
dependent properties in the sense that they are reciprocally producers and
products, means and ends, in a network of synthetic pathways. […] But in this
case, this hierarchic ontological dependency is tangled in what Douglas Hof-
stader has called ‘strange loops’». T. D EACON , Incomplete nature: How mind
emerged from matter, p. 178.
96 Victor Marques & Carlos Brito Articles

cling-around, a succession of shifts in level that feels like an


upward movement in a hierarchy turns out to give rise to a closed
cycle: in violation of the seeming hierarchy, one winds up where
one started33. That is exactly what happens when Rosen creates
an impredicativity in order to avoid the infinite regress, by
folding back the hierarchy into a loop 34. Varela, as well as
Rosen, stresses the necessity of both hierarchy and circularity,
resulting in a seemingly paradoxical structure that defies atomic
analysis35. This type of system seems to resist modeling by
conventional mathematical tools. Rosen (1991) uses Category
Theory in order to accommodate impredicative loops in a math-
ematical formalism, Varela (1975) uses Spencer Brown’s calculus
of indications to develop a «calculus for self-reference»36, and,
more recently, Chemero and Turvey (2008) suggested hyperset
theory as a tool to model autonomy and complex systems with

33 D. HOFSTADTER, What is it like to be a strange loop?, in Self-representational


Approaches to Consciousness, ed. U. Kiregel and K. Williford, MIT Press, Cam-
bridge 2006, p. 465.
34 Cf. R. ROSEN, Essays on Life Itself. Columbia University Press, New York

2000; S.W. KERCEL, The Endogenous Brain, «Journal of Theoretical Biology »


LIV (3-1), 2004, p. 61; J-H.S. HOFMEYR, The biochemical factory that autonomously
fabricates itself: A systems-biological view of the living cell, in Systems Biology: Philosophical
Foundations, ed. F.C. Boogerd, F. Bruggeman, J.-H.S. Hofmeyr and N.V. Wester-
hoff. Elsevier, Amsterdam 2007, p. 217.
35 Varela attributes the reluctance to concede a central role to circularity to the

influence of what he calls «Fregean viewpoint», which he associates with the


theory of types in logicism and the atomism of well-founded sets theory: «The
mental picture is that of a tree with roots and branches. But this view is
awkward for describing whole systems, where the picture is more that of a
closed network with roots and branches intertwining […]. No type distinctions
are possible in such a network». F. VARELA, Principles of biological autonomy ,
p. 167. Deacon et al also relate the circularity of life with circularity in the
context of linguistic or formal systems: «this is analogue of self-reference, a
logical type violation, and it is not surprising that this feature is even the
defining characteristic of reflexive reference in language».
T. DEACON, J. HAAG & J. OGILVY, The Emergence of Self. in In Search of Self:
Interdisciplinary Perspective on Personhood, ed. J.W. van Huyssteen & E.P. Wiebe.
William B. Eerdmans Publishing Co., Gran Rapids 2010, p. 333.
36 F. VARELA, A Calculus for self-reference, «Int. J. General Systems» LIV (2), 1975, p. 5.
Articles Rise and Fall of the Machine Metaphor 97

circular organization37. The contrast between Maturana and


Varela, on one hand, and Rosen is that the later puts a lot more
focus on the internal functional differentiation of the system itself
and emphasizes that we are dealing not just with a mere material
cycle, but a circularity in terms of control conditions of metabolic
transformations. Whereas the autopoietic model is more focused in
the spatial enclosure with membranes, Rosen is more preoccupied
with the logical organization, expressed in the formal language of
category theory.
Rosen, however, is keen to draw attention to his conclusion
that the relational description of machines does not allow causal
closure, and the circular description is at the very core of what
makes life distinct from non-life. So at least for the late Rosen, an
organism is not a machine at all. Although Matura and Varela call
their systems «autopoietic machines», in order to explicitly fulfill a
«non-animistic criteria», they are so different from conventional
machines in their autonomy, plasticity and circular organization
that it is possible to question to what extent the use of the term is
really useful or adequate.
The similarities between Rosen’s and Varela’s models are
now amply recognized, by Roseans and Vareleans alike38. But

37
For Chemero & Turvey what all those models, including Kauffman ‘s auto-
catalytic sets, share is «having every function be a product of a system, having
loopy hyperset diagrams that terminate only with raw materials». They are
closed to efficient cause and open to material cause. A. CHEMERO & M.
TURVEY, Autonomy and hypersets, «Biosystems» LIV (91), 2008, p. 320.
38 Kercel, a Rosean, states: «In processes of life and mind, Rosenesque com-

plexity if equivalent to autopoiesis. Its distinguishing feature is a hierarchical


closed-loop of causal entailment». S.W. KERCEL, The Endogenous Brain. «Journal
of Theoretical Biology» LIV (3-1), 2004, p. 61. But also Evan Thompson, a
long time Varela ‘s collaborator: «Like Maturana and Varela, Rosen aims to give
a precise account of the organization of life, and although they never mention
each other in their writings, there are deep affinities between their theories.
Unlike Maturana and Varela, however, Rosen presents a rigorous argument for
distinguishing between organisms and machines. An intriguing feature of this
argument is that it is precisely what Maturana and Varela would call the circular
and self-referential organization of the living that distinguishes organisms
from mechanisms and machines». E. THOMPSON, Mind in Life: Biology,
Phenomenology, and the Sciences of Mind, Belknap Press, Cambridge 2007, p. 141.
98 Victor Marques & Carlos Brito Articles

both approaches, being purely relational, emphasize organization-


al demands while paying almost no attention to material and
thermodynamical conditions. Autopoiesis was explicitly con-
ceived as an abstract machine, independent of the nature of the
materials39, and Rosen’s diagrams convey the loopy form of those
systems, but say nothing about the nature of constraints and
chemical process that are actually able to realize this distinct
form40. That is not the case with the more recent efforts of
Kauffman, Mossio & Moreno and Deacon, where the emphasis
also lays on the far-from-equilibrium and thermodynamically
open nature of this kind of system. Being interested not only on
synchronic analysis, but also in the actual historical genesis of
autonomy within the material world, and struggling to give a
diachronic account of how material systems came to express this
peculiar circular organization, they offer a bottom up approach,
which builds from physical self-organizing process up to biologi-
cal, and cognitive, complexity.
Kauffman’s effort is to give a naturalized account of agency,
explaining how can autonomous systems, that «act on their own
behalf», exist in nature and evolve from inanimate matter. Noting
that constraint and work are related terms, since work is con-
strained release of energy and in general (although not always) it
takes work to build constraints, Kauffman defines organism as an
autocatalytic set that does at least one work-constraint cycle.
Living beings thus couple spontaneous and nonspontaneous
processes to build constraint on the release of energy and use the
resulting work to reproduce their own constraints.
The basic idea of a work-constraint cycle is also found in
Deacon, who, following Kauffman, stresses that «besides being a

39 «For Maturana and Varela, autonomous systems are defined by the abstract
property of operational closure, leaving aside material and energetic require-
ments». X. BARANDIARAN & A. MORENO, Adaptivity: From Metabolism to Behav-
ior, « Adaptive Behavior» LIV (16-5), 2008, p. 325.
40
As Chemero & Turvey observes: «From our point of view, Rosen’s metabolism-
repair systems and Maturana and Varela’s autopoietic systems are valuable as
characterizations of an abstract property, while Kauffman’s work is valuable for
connecting this abstract property to chemical processes». A. CHEMERO & M.
TURVEY, Autonomy and hypersets. «Biosystems», LIV (91), 2008, p. 320.
Articles Rise and Fall of the Machine Metaphor 99

product of work, constraint is also a precondition of work», and


that «the maintenance, reconstruction and reproduction of dynam-
ical constraints is a core characteristic of life»41. What they all share
is the insight of organisms as dissipative systems and the emphasis
on a kind of continuity between the simple self-maintenance of
physical self-organizing processes, characterized by the spontane-
ous emergence of dynamical, precarious constraints, and life
proper, with its enduring, propagational logic. The circular form
must be realized in far from equilibrium conditions, which imply
that organisms, by being organizationally closed, are also neces-
sarily thermodynamically open.
Although it is true that both machines and organisms are
fundamentally characterized by their organization, whose physical
realization involves material structures that constrain underlying
processes, the nature of these structures differs profoundly from
one case to the other. While the machine components are often
rigid and inert, in organisms they are typically precarious , mallea-
ble, plastic. The order and stability expressed by organic matter in
its form do not result from stable materials, in contrast to ma-
chines, where the constraints are built to resist deformation and
to be relatively insensitive to thermodynamical fluctuation. Ra-
ther, the organic dynamics is based on non-equilibrium chemical
processes, generating regularity and order by self-organization
and self-assembly. Organisms are dynamical systems out of equi-
librium – as dissipative systems they are constantly renewing their
material composition and requiring energy input to maintain their
constraints. In order to produce order spontaneously, they must
be thermodynamically open42.

41 T. DEACON, Incomplete nature: How mind emerged from matter, cit., p. 262.
42 «This property of causal closure in ‘soft material automata ‘ (as opposed to
the rigid or fixed structure of relationships in traditional man-made machines)
involves high rates of energy dissipation, so it requires the continuous produc-
tion of work by the system. Thus, living systems, which are continuously and
literally fabricating themselves, can only maintain their organization in far from
equilibrium conditions by being material-thermodynamically open». A. MORE-
NO, K. RUIZ-MIRAZO, X BARANDIARAN, The impact of the paradigm of complexity
on the foundational frameworks of biology and cognitive science,i Handbook of the Philoso-
100 Victor Marques & Carlos Brito Articles

Terrence Deacon rightly calls attention to the fact that


what distinguishes organisms from other dissipative systems is
that in the case of living beings there is not merely a process of
self-organization, but rather «a reflexively organized constella-
tion of self-organizing processes»43. The self-organization
characteristic of the organism as a whole is distinctively of
second order kind, a meta-self-organization: «The constraint
maintaining propagation logic of the organism is in a sense a
higher-order self-organizing dynamic among component self-
organizing processes»44. For Deacon, it is the synergetic ar-
rangement of self-assembly and non-equilibrium self-
organizing processes that gives rise to the «teleodynamical
features» of living beings, their «immanent purposiveness». The
correlated co-production of constraints ensures the perpetua-
tion of this holistic co-dependency, resulting in what Deacon
calls (making reference to Varela) a «virtual self»45.
Due to their precarious nature, biological structures are al-
ways being built and degrading, and a functional component
persists not because it is structurally rigid, but rather because it is
incessantly produced and reproduced by the collective activity of
the other components. It is at this level of reciprocal constraint
generation that we find the unique circular pattern of life. Materi-
al structures that play the functional role of organizational con-
straints reinforce and remake each other, thereby maintaining the
identity of the system as a whole, in what Mossio & Moreno
denominated «organizational closure»:

The main idea is that biological systems are able to maintain


themselves by constituting a web of structures exerting mutual
constraining actions on their boundary conditions, such that
the whole web is collectively self-maintaining. The mutual de-

phy of Science. Volume 10: Philosophy of Complex Systems, ed. C. Hooker. North
Holland Elsevier, Amsterdam 2009, p. 325.
43 T. DEACON, J. HAAG & J. OGILVY, The Emergence of Self, in In Search of Self:

Interdisciplinary Perspective on Personhood, ed. J. W. van Huyssteen & E. P. Wiebe.


William B. Eerdmans Publishing Co., Gran Rapids 2010, p. 319.
44 Ibidem.
45 T. DEACON, Incomplete nature: How mind emerged from matter, cit., p. 311.
Articles Rise and Fall of the Machine Metaphor 101

pendence between a set of constraints is what we call organiza-


tional closure46.

Being built on soft constraints, organisms have to be organi-


zationally closed to persist through time for longer than their
constituent components, and this organizational closure in turn
demands that they are necessarily material-thermodynamically
open systems, engaged with the environment in order to feed
their own process of self-fabrication in far from equilibrium
conditions. Machines, on the other hand, do not express organi-
zational closure. Machines don’t have a metabolism, and their
parts are not produced internally, each component being fabricat-
ed separately and arranged during the assembling process. The
organic components, by contrast, are integrated and interdepend-
ent from the outset. The rigid nature of machine components
give them a self-subsistent aspect, they can be easily separated
from the rest of the machine, and the process of analysis, of
dividing a machine into pieces, doesn’t modify at all their essen-
tial nature: they remain basically the same whether within or
without the rest of the system. However, in the case of organ-
isms, components tend to cease to exist, in a sense they lose their
“essence”, when removed from the context of the overall organic
activity, within which they are continuously being reproduced and
repaired. While machines are assembled from pre-existing com-
ponents that have independent existence, in living beings the
existence of each component depends on the context afforded by
the other components, in a tangled web of mutual dependence
and reciprocal causation.

4. Natural purposes

As several authors have recently called attention, this view of


life based on the idea of closure resembles the Kantian notion of
“natural purpose”. In the second half of the third critique, reflect-

46M. M OSSIO & A. M ORENO, Organisational Closure in Biological Organisms,


«Hist. Phil. Life Sci.», LIV (32), 2010, p. 269.
102 Victor Marques & Carlos Brito Articles

ing about the question of purposiveness in nature, Kant draws a


distinction between external, relative, teleology, that which we
find in mechanical artifacts, and internal, intrinsic, teleology47.
Organisms, for Kant, are organized and self-organizing beings,
both causes and effects of themselves.
Following that line of thought Kauffman coined the term
«Kantian wholes» to refer to such natural systems as organisms,
where the whole exists for and by means of the parts and the
parts for and by means of the whole48. As Juarrero highlights,
reclaiming Kant’s original terminology for contemporary uses
(specifically for dealing with complex systems, offering a natural-
istic account of agency), a «natural purpose» is an object in which

a member is not only a means but also an end; it both contrib-


utes to the whole and is defined by it. No machine exhibits this
kind of organization, for the efficient cause of a machine lies
‘outside’ the machine in its designer, and its parts do not owe
their existence to each other or to the whole49.

In the last two and a half decades, biologists and philoso-


phers of biology have shown an increasing interest in the Kantian
approach to organisms, reigniting a philosophical discussion
about the concept of natural purpose and its place in a naturalist

47 «In such a product of nature each part is conceived as if it exists only


through all the others, thus as if existing for the sake of the others and on
account of the whole, i.e., as an instrument (organ), which is, however, not
sufficient (for it could also be an instrument of art, and thus represented as
possible at all only as an end); rather it must be thought of as an organ that
produces the other parts (consequently each produces the others reciprocally),
which cannot be the case in any instrument of art, but only of nature, which
provides all the matter for instruments (even those of art): only then and on
that account can such a product, as an organized and self-organizing being, be
called a natural end».
I. KANT, Critique of the Power of Judgment, tr. P. Guyer & E. Matthews, Cam-
bridge University Press, Cambridge 2000, p. 245.
48 See, for instance: G. LONGO, M. MONTÉVIL, S KAUFFMAN, No entailing laws,

but enablement in the evolution of the biosphere. arXiv:1201.2069 2012.


49 A. JUARRERO, Dynamics in Action. Intentional Behavior as a Complex System, MIT

Press, Cambridge 1999, p. 47.


Articles Rise and Fall of the Machine Metaphor 103

project. Juarrero (1999)50, Weber & Varela (2002)51, Moss


(2003)52, Thompson (2007)53, Kauffman (2000)54, Moreno &
Mossio (2010)55 and Deacon (2012)56 – all refer back to Kant and
his pioneering use of self-organization as conceptual tool to
understand purposiveness in natural systems and what distin-
guishes organisms from artificial machines. What unites all those
authors is the shared intellectual aim to naturalize, rather than
simply eliminate, teleology.
As we have seen, modern thought, in order to understand
living beings as fully natural parts of a thoroughly non-
teleological physical reality, initially replaced the intrinsic purpose
of Aristotelian type by the external purpose derived from a trans-
cendent creator. The tendency to see the teleology of organisms
as merely relative, derived, not intrinsic to the organism itself,
persists even with the emergence of evolutionary thought, to the
extent that the process of natural selection is interpreted as anal-
ogous to the action of an intelligent designer (as in Dennett and
Dawkins, for example). Kant, in contrast, proposes a conception
of the organism as a self-organizing entity, in general lines quite
similar to the contemporary theory of autopoiesis – as Varela
himself eventually recognized57.
Kant, however, found himself faced with the dilemma,
named by him as the «antinomy of teleological judgment», be-

50 A. JUARRERO, Dynamics in Action. Intentional Behavior as a Complex System, MIT


Press, Cambridge 1999.
51 A. WEBER & F. VARELA, Life after Kant: Natural ends and the autopoietic founda-

tions of biological individtuality. «Phenomenology and the Cognitive Sciences»,


LIV (1), 2002, p. 97.
52 L. MOSS, What Genes Can’t Do, MIT Press, Cambridge 2004.
53 E. THOMPSON, Mind in Life: Biology, Phenomenology, and the Sciences of Mind.

Belknap Press, Cambridge 2007.


54 S. A. KAUFFMAN, Investigations, Oxford University Press, Oxford 2001.
55 M. M OSSIO & A. M ORENO, Organisational Closure in Biological Organisms.

«Hist. Phil. Life Sci.», LIV (32), 2010, p. 269.


56 T. D EACON , Incomplete nature: How mind emerged from matter, Norton

& Company, New York 2012.


57 A. WEBER & F. VARELA, Life after Kant: Natural ends and the autopoietic founda-

tions of biological individtuality, «Phenomenology and the Cognitive Sciences» LIV


(1), 2002, p. 97.
104 Victor Marques & Carlos Brito Articles

tween the demands of Newtonian physics, according to which all


natural objects should be understood in purely mechanistic terms,
which are integral to Kant’s own conception of nature, and our
experience in dealing with the biological realm, that seems to
force us to think in terms of final causes. Kant himself was quite
pessimistic regarding the possibility of the human mind to explain
«organized beings» (natural purposes) according to «mere me-
chanical principles of nature».
On the one hand, the phenomena of life lead us to think the
idea of a «natural purpose», in which the parts are only possible
by reference to the whole, and whose organization has nothing
analogous in the rest of nature:

An organized being is thus not a mere machine, for that has on-
ly a motive power, while the organized being possesses in itself
a formative power, and indeed one that it communicates to the
matter, which does not have it (it organizes the latter): thus it
has a self-propagating formative power, which cannot be ex-
plained through the capacity for movement alone (that is,
mechanism)58.

On the other hand, the very idea of an «organized and self-


organizing» being, that is «both cause and effect of itself», is, at
least according to Kant, beyond comprehensibility. For Kant, the
very idea of a material vitality, implied in a realist, ontological
interpretation of natural purpose is in fact a non-sense, since it
contradicts the very essence of matter:

the possibility of a living matter (the concept of which contains


a contradiction, because lifelessness, inertia, constitutes its essen-
tial characteristic), cannot even be conceived59.

The dilemma is rooted in the utter incompatibility between a


concept of life that involves intrinsic teleology and a concept of
nature that does not allow for the self-organization of matter.
However, as stated by Thompson60:

58 I. KANT, Critique of the Power of Judgment, cit., p. 246.


59 Ivi, p. 265.
60 E. THOMPSON, Mind in Life, cit., p. 140.
Articles Rise and Fall of the Machine Metaphor 105

This dilemma no longer seems compelling. Our conception of


matter as essentially equivalent to energy and as having the po-
tential for self-organization at numerous spatiotemporal scales is
far from classical Newtonian worldview. In particular, the phys-
ics of thermodynamically open systems combined with the
chemistry and biology of self-organizing systems provides an-
other option that is not available to Kant: life is an emergent
order of nature that results from certain morphodynamical prin-
ciples, specifically those of autopoiesis.

Autopoiesis theory thus offers a contemporary naturalized


version of the Kantian notion of natural purpose 61. The notion of

61 Although we don’t have the space to develop this point further here, we
would like to at least remark briefly that, more than Kantian, this position
seems to have a distinctive ʻGerman Idealismʼ flavor to it. Hegel in his philos-
ophy of nature, while cheering Kant for the resumption of the notion of
internal teleology, already present in Aristotle but lost in modern philosophy, is
also a critic of what he sees as ambivalence and hesitation in the Kantian
position. With the current growing interest in the Kantian concept of
Naturzweck the Hegelian critique of Kant’s hesitation becomes relevant and
surprisingly contemporary. Building upon Kant, Hegel offered a conceptual-
ization of life based on «inner purposiveness», «assimilation», «self-referring
negative unity», and self-determination as self-limiting – a philosophical
account of living beings as incomplete wholes. But unlike Kant, Hegel never
understood this circular organization (being both cause and effect of itself) as
beyond comprehensibility, nor saw any incompatibility between life and matter.
In his Philosophy of Nature Hegel even identified life with chemical process
circularly arranged. In fact, some biologists and philosophers who, like
Thompson, try to rescue Kant’s third critique end up looking more like Hegel,
in the sense that what they really want is not only a mere regulative principle,
but the proper German idealist insight that the Kantian notion of life can gain
an ontological interpretation and life can thus serve as a starting point for
questioning the traditional mechanical image of nature. What Kauffman refers
as «Kantian whole» can thus more properly be called «Dialectical wholes». This
view of life as natural end, a metabolic and circular conception, was basically
inherited by the philosophy of dialectical materialism, which in turn had a
concrete impact, via the work of Marxist (or quasi-marxist) biologists, usually
the main critics of the machine metaphor and very much engaged in the
development of theoretical biology; one just has to think of J.B.S. Haldane,
Joseph Needham, J.D. Bernal, Oparin, Conrad Waddington, Richard Levins
and Richard Lewontin. The considerable influence of Marxist ideas on the
anti-reductionist camp of biological science is a story still waiting to be told.
106 Victor Marques & Carlos Brito Articles

organizational closure allows us to explain the continued exist-


ence of a system as a function of its own activity. The function of
a particular component is found on its particular contribution to
the persistence of a totality that produces it and on which it
depends and relies for its own existence. Accordingly, purpose
may be naturalized without reference to an external artifice, not
even an analogy with the process of conscious design. Indeed,
authors such as Varela actually put in question the analogy be-
tween evolution and engineering and the popular image of natural
selection as a kind of designer (a reified «mother nature»), even if
a blind one. Reversing the path of modernity, Varela rediscovers
naturalistic teleological thinking, making purpose immanent once
again in the real material system circularly arranged, and negating
altogether the external teleology of a transcendent designer as a
good model to think biology.
Contemporary science went beyond the vision of matter as es-
sentially inert and passive, developing at least the foundations of a
science of self-organized complex systems. In particular, a thermo-
dynamic theory of organizational closure seems to accommodate
all the determinations of «natural purposes», in so far as it charac-
terizes life as a network of processes which produces its own
components and also the topological boundary that defines it as an
individual entity: being and doing are one and the same process in
this self-reproducing dynamics. In such a network the parts mutu-
ally produce each other precisely in accordance with Kant’s defini-
tion of natural purpose, so the whole network can be viewed as
cause and effect of itself. The result is a self-perpetuating totality
that emerges out of local processes, but is at the same time the
condition, the presupposition of these processes.
By moving in that direction, however, we seem to be back to
something similar to the naturalism with final cause so dear to
Aristotelian thought. We are thus forced to recognize that Aristo-
tle was not completely wrong, and that his defense of teleology
inherent in the material world, particularly in respect to organized
and self-organizing beings, that is to say, organisms, is not absurd
or inconsistent with an updated and enriched form of naturalism.
Articles Rise and Fall of the Machine Metaphor 107

As noted by Moss62:

A sense of similarity between Aristotle’s hylomorphic under-


standing of soul and much more recent descripitions of self-
organization dynamic systems is not entirely accidental. […] the
idea that epigenesist was achieved by self-organizing movements
driven by internal orientation toward an adapted form was en-
tirely consistent with his metaphysics[…] Aristotle, by contrast,
and epigenesists ever since, have endeavored to explain life-
forms not as artifacts designed from without but as self-
organizing, ‘autopoietic’, ends-unto-themselves.

It is worth noting, however, that nothing of what was dis-


cussed above requires us as theoretical commitment to any com-
prehensive teleology. Just as the primitive mechanicism of modern
physics sought to impose on the entire universe an impoverished
notion of matter, extrapolated from the observation of simple
systems, Aristotelian hylozoism extrapolated in the opposite direc-
tion, generalizing the particular experience of living beings to the
cosmos as a whole. But the fact that teleology is a natural reality,
that is to say, is objectively real in the world, does not mean that it
must be the reality of the whole of nature. That some material
objects allow naturally explanation by final causality does not imply
that all natural objects have a final cause. Natural purposes, Kanti-
an wholes, are quite particular types of material systems that
emerge from a specific organization of physical-chemical process,
and it is completely plausible that this type of organization is only
manifest in a very limited range of phenomena.

5. Conclusions

The perception that living beings are something distinctively


different from inert matter and the attempt to explain this differ-
ence in scientific terms has been a constant theme in the history
of the human inquiry on nature. One possibility consisted in
postulating a difference in substance: life would be the manifesta-

62 L. MOSS, What Genes Can’t Do, cit., p. 8.


108 Victor Marques & Carlos Brito Articles

tion of a special sort of substance, the so-called elan vital, which is


absent in ordinary physical objects, and the distinctive features of
the living were to be explained in terms of the special properties
of this substance. Eventually, technology advanced to the point
where the material constitution of both living and inanimate
matter could be examined in minute detail, and no difference in
substance was ever found. That same investigation, however,
revealed living beings as extremely complex systems, composed
of heterogeneous parts, each seeming to express a definite pur-
pose and involved in very intricate relationships with other parts.
So, the difference should be in the arrangement: living beings are
organized physical systems, as opposed to sand and rock which,
even though composed by heterogeneous materials, cannot be
said to be organized. But manmade machines are also organized
systems, which can be appropriately described in terms of com-
ponents with specific functions and relationships among those
components. Of course, men have never been able to construct a
machine which resembled, not even closely, the complexity of the
simplest living organism. In spite of that, the machine paradigm
eventually became dominant, establishing as a general conviction
that there is no essential difference between machines and organ-
isms, a mere difference in degree of complexity, which is to be
inevitably reduced and perhaps eliminated as technology advances.
The quest for the ultimate facts that would separate living
beings from inanimate matter was eventually replaced by the
quest for the ultimate facts about the organism as an organized
physical system, under the machine paradigm. The machine
should be taken not only as a metaphor, but as a model for the
living. That movement inaugurated what could be called an ana-
lytical research program in biology: the living organisms were to
be understood in terms of components parts with independently
assessed properties, and relationships among those parts. This
strategy was extremely successful as witnessed even today by the
increasingly fast development of molecular biology, and the
progressive identification of ever more specific components and
processes involved in metabolism. The main goal in this line of
research is to achieve a complete understanding of the physico-
chemical operation of a living organism, in the same way that the
Articles Rise and Fall of the Machine Metaphor 109

operation of a machine can be completely understood by a thor-


ough examination of all its parts.
Maturana and Varela, and Robert Rosen, begin all their dis-
cussions by emphasizing that what is most relevant about the
living organism is to be found in its organization. More precisely,
they first claim that in order to understand the phenomenon of
life biology should adopt the same type of relational description
already applied to machines, where components are presented in
terms of the abstract functions they support, and the system as a
whole is described in terms of a set of relations among those
components. They go so far as to postulate a principle of inde-
pendence between organization and physical structure: specific
material details are unimportant in the sense that several material
structures can typically instantiate the same properties and func-
tions that characterize the components. As a natural consequence,
they propose to that what is special about life cannot be found by
examining the particular composition and properties of its material
structure, but rather by attempting to have a glimpse of what is
essential about its overall organization. It is in this context that they
indicate that the specificity of living organisms resides on the
special type of relations in which their components are involved:
relations of mutual production and interdependence. Every com-
ponent in a living organism has a natural tendency to decay, but
the long lived existence of the system is made possible by a perma-
nent activity of self-maintenance or self-production, where com-
ponents repair and produce other components. Their position can
be restated in a few words as follows: not operation but construc-
tion is the key to understand life. Those very same ideas can also
be found in the work of Robert Rosen.
Now, it seems to us that the fact that living beings must be
explained in terms of processes of material construction does
intertwine organization and material structure again. If correct,
this observation would have the consequence that a purely analyt-
ical approach to biology is not viable. Living beings are entities
which only exist as such while embodied as material assemblage
of interdependent components. In a situation where the existence
of those components is a given, that is, their construction and
disintegration lie outside the scope of the dynamics under investi-
110 Victor Marques & Carlos Brito Articles

gation (e.g., when we are trying to understand the operation of a


machine), abstracting away materiality is not a real issue and the
analytical approach is unproblematic. But, in the case where the
components are not stable entities, being constantly subjected to
processes of construction and disintegration, it is not clear any-
more if they are the sort of thing that can be properly described
as standing in fixed, well-defined relationships with each other.
Here, materiality must be taken into account because, since the
components are composite material structures, their specific
properties and functions, and the relations they can participate in,
are determined by their specific material form which, in turn, is
the result of the processes of material construction which consti-
tute the activity or operation of the living organism. Now we can
see more clearly the challenge posed by the living organism to the
traditional mode of analytical explanation. The basic and tacit
assumption of this method is that we can always explain the
operation of a system in terms of properties and relations among
well-defined components. But, in the case of the living organism,
these very properties and relations are the (indirect) result of the
operation of the system, and so we cannot take the first for
granted in order to explain the later. In other words, we cannot
find the basic elements which are required to support the analyti-
cal explanation.
Any adequate explanation of the life process will necessarily
have to consider the living organism as an organized system of
material aggregates, whose specific properties and functions
depend on their particular form, which are subject to disintegra-
tion but are constantly repaired and replaced by the activity of
material construction of the system. Moreover, if this is indeed
the case, the analytic method is not enough if we want to under-
stand life. This, of course, does not imply that the analytic meth-
od is of no use in biology. Mechanical explanations can be used
unproblematically to explain the operation of specific metabolic
processes (e.g. the citric acid cycle) by assuming the existence of
stable components and relations among them. We only have to
give up the idea that if we obtain sufficient explanations of this
kind we can put all of them together to have an explanation of
life itself. Last, but not least, we emphasize that the abandonment
Articles Rise and Fall of the Machine Metaphor 111

of the machine metaphor in anything weakens naturalist com-


mitments, and actually may be a crucial step to truly enrich our
conception of nature and open the path towards a fully, and
compelling, naturalization of teleology and normativity, with
important consequences to the study of evolutionary processes
and cognition. As Lewontin & Levins once argued63:

The program of Harvey and Descartes to reveal the details of


the bête machine has worked. The problem is that the machine
metaphor leaves something out, and naive mechanistic biology,
which is nothing physics carried on by other means, has tried to
cram it all in at the expense of a true picture of nature.

63R. LEWONTIN & R. LEVINS, Biology under the influence. Dialectical essays on ecology,
agriculture, and health, cit., p. 222.
BLUMENBACH ON TELEOLOGY AND THE LAWS OF
VITAL ORGANIZATION

by François Duchesneau

Abstract. Blumenbach’s physiology and natural history are grounded on a


complex theory of vital forces. These forces correspond to a hierarchy of dy-
namic principles that include the Bildunsgstrieb as the proper sufficient reason
for sequences of epigenetic processes. From an epistemic viewpoint, these
dynamic principles are presumed to stand for the concealed causes of constant
and regular physiological effects. But in light of Blumenbach’s methodology
does not such a system of specific forces point to the project of formulating laws
of vital organization that would be ultimately irreducible to physical and
chemical laws? Presuming that such is the case, what kind of constitutive or
regulative use of teleology was supposed to fit into the formulation and justifi-
cation of those laws? Assessing the methodological and epistemic profile of
Blumenbachian physiology might throw some light on the experimental
pattern of Entwicklungsgeschichte that became a prominent feature of early
19th century German biology.

Johann Friedrich Blumenbach (1752-1840) has often been


considered as the principal initiator of biological vitalism in Ger-
many at the turn of the 19th century1. This is partly due to the
fact that he managed to graft a hypothesis on the formation of
organisms by epigenesis to an analytic physiology in the style of

1 T. LENOIR, The Strategy of Life. Teleology and Mechanics in Nineteenth Century


German Biology, Reidel, Dordrecht, 1982, was quite right in pointing to the
influence of Blumenbach’s physiology on later German biological theories. It
is less evident that, according to that interpretation, Blumenbach developed
directly similar views to Kant’s and professed a coherent and systematic «vital
materialism». Lenoir’s interpretation has been notably challenged by R.J.
RICHARDS, Kant and Blumenbach on the Bildungstrieb: a historical misunderstanding,
«Studies in the History and Philosophy of Biological and Biomedical Sciences»,
XXXI, 2000, pp. 11-32. For a recent appraisal of this controversy, see J.H.
ZAMMITO, The Lenoir Thesis Revisited: Blumenbach and Kant, «Studies in the
History and Philosophy of Biological and Biomedical Sciences», XLIII, 2012,
pp. 120-132.
Verifiche XLIII (1-4), 2014, pp. 113-135.
114 François Duchesneau Articles

Haller. This hypothesis attributes the processes of generation to a


formative drive that he terms ‘Bildungstrieb’ or ‘nisus formativus’.
This theoretical entity, apart from meaning a principle of vital
self-organization, was further designed to provide a sufficient
reason, a methodological key, for explaining the regular processes
that unfold in integrative fashion through the various functions
of organic parts. Did the reliance on such a hegemonic principle
and on the vital forces which were presumed to derive from it
mean that it was impossible to ‘mechanically’ explain the process-
es by which organisms frame up, operate, self-preserve, self-
repair and reproduce? Or did it mean that one could not explain
living organisms and their operations otherwise than by resorting
to teleological concepts? Were the Bildungstrieb and the other vital
forces only substitutes for presumed, yet unknown or unknowa-
ble, ‘mechanisms’? Or did they represent laws of vital organiza-
tion that it would have been vain to try to reduce to orderly
causal sequences of physical and chemical operations? These
questions point to methodological antinomies that have played a
major role in subsequent biological theorizing.
I shall try to show what particular methodology of investiga-
tion, explanation and justification was at work in Blumenbach’s
most influential theory of vital forces, especially in view of the
fact that the Bildungstrieb hypothesis shifted from being an ac-
count of what happens to individual organisms in terms of physi-
ological processes to becoming an account of what is involved in
the transformation of species as breeds of organisms that vary
and diversify in time: thus Blumenbach enlarged the domain of
application for his type of vitalist scientific approach. But he
fostered a significant posterity who remodelled his physiological
approach in diverse ways and raised variants of it up to the more
recent period. I shall here limit myself to considering the epige-
netic theory that Blumenbach developed in Ueber den Bildungstrieb
(1st edition 1781; 2nd and 3rd 1789 and 1791) and in connected
writings, such as De nisu formativo et generationis negotio nuperæ observa-
tiones (1787). This doctrine has to be interpreted within the
framework of the theory of vital forces that Blumenbach exposed
in his Institutiones physiologicæ (1st edition 1787) and Handbuch der
Naturgeschichte (post-1781 editions). Furthermore, in the later
Articles Blumenbach on Teleology 115

versions of his De generis humani varietate nativa (originally pub-


lished in 1776) Blumenbach transposed his notions of Bildungstrieb
and vital forces from the biological consideration of individual
organisms to that of species, viewed as phyla of kindred organ-
isms, subject to variations: hence an enlarged theoretical context
for the development of his so-called vitalist methodology; but I
shall not for the present deal with this aspect of his theory2.

1. Epigenesis and Bildungstrieb

«Causa latet, vis est notissima» – «The cause is hidden, but the
force is most evident»): it is by means of this quotation from
Ovid’s Metamorphoses3 that Blumenbach relates his notion of
Bildungstrieb to the category of Newtonian explicative causal prin-
ciples4. Like attraction, the nisus formativus signifies a natural pow-
er, a force whose determination rests in the constant and regular
empirical effects to which it relates, even if the cause of these
effects remains concealed, as would be the case with ‘occult
qualities’5. What delimits the field of phenomena for which the
nisus stands as a sufficient reason is a sequence of epigenetic
phenomena that characterize the transition from an amorphous
organic matter at the inception of developmental processes to the
functional architecture of the resulting organism at its various
stages of development. The phenomena of generation proper, but
also those of growth, nutrition and reproduction, are referred to
the formative nisus6. This vital force expresses the constancy and

2 For a general analysis of Blumenbach’s physiology, see F. DUCHESNEAU, La


Physiologie des Lumières. Empirisme, modèles et théories, Garnier, Paris, 2012, pp. 492-520.
3 OVID, Metamorphoses, IV, v. 287.
4 It is still useful to refer here to T.S. H ALL, On Biological Analogs of Newtonian

Paradigms, «Philosophy of Science», XXXV, 1968, pp. 6-27.


5 J.F. BLUMENBACH, Über den Bildungstrieb, Johann Christian Dieterich, Göttin-

gen, 1791, p. 33.


6 Ivi, p. 32: «Ein Trieb, der folglich zu den Lebenskräften gehört, der aber eben

so deutlich von den übrigen Arten der Lebenskraft der organisirten Körper
(der Contractilität, Irritabilität, Sensilität etc) als von den allgemeinen phy-
sischen Kräften der Körper überhaupt, verschieden ist; der die erste wichtigste
116 François Duchesneau Articles

generality, that is, the lawfulness of the processes of organic


formation as revealed by experience.
To establish the specific order that characterizes generative
processes, Blumenbach follows two paths. One consists in identi-
fying the sequence of phenomena resulting from the mixing of
seminal fluids, that is, the upgrading of organic components to a
level allowing for the emergence of a Bildungstrieb, then in identi-
fying the formative stages of embryos and membranes as they
progressively frame up. Along the other, more analogical, path,
Blumenbach intends to show that the generative and formative
force follows from an extension of the organizational disposition
to be found in inanimate beings, which disposition is accounted
for without resorting to pre-existent germs. If the nisus properly
belongs to organic configurations, it prolongs in a way a propen-
sity to organize and get organized that shows up throughout the
whole extent of ‘physical phenomena’7. The processes of tissue
nutrition reveal the intervention of such a productive force while
showing the specificity of the generative order compared to any
mode of inorganic formation8.
Blumenbach’s theses on nutrition are exposed for instance in
his Institutiones physiologicae. Blumenbach surmises that the solid
parts of the organism which partake in nutrition have all of them,
correlatively and to various degrees, the faculty of self-

Kraft zu aller Zeugung, Ernährung, und Reproduction zu seyn scheint, und


den man um ihn von andern Lebenskräften zu unterscheiden, mit dem Namen
des Bildungstriebes (nisus formativus) bezeichnen kan».
7 J.F. BLUMENBACH, De nisu formativo et generationis negotio nuperæ observationes,

Apud J.C. Dieterich, Gottingæ, 1787, pp. xi-xii: «Etiamsi enim facile largiamur
ingens adhuc intercedere discrimen haecce inter mineralia et organica e con-
trario corpora : ex eo tamen capite ad nostram quam agimus disquisitionem
utilia esse possunt, quod determinatarum formarum exempla exhibent, in
quibus de germinibus eiusdem formae præexistentibus, neminem serio res
agentem cogitare quidem licet; idque eo minus cum idem aurichalcum, ex
iisdem quibus constat metallis, pro diverso parandi modo diversas quoque et
toto cœlo ab se invicem differentes crystallorum formas exhibeat; id quod
primo statim intuitu comparando eam quam diximus speciem cum altera quam
Stück-Messing vocant, apparet».
8 Caspar Friedrich Wolff had steered a discussion with Blumenbach on that

ground: hence Blumenbach’s essay, Ueber die Nutritionskrafl (1789) for the
competition launched by the St. Petersburg Academy on that topic.
Articles Blumenbach on Teleology 117

reproduction. At a lower level, this merely means the maintaining


of the organic integrity through continuous and imperceptible
changes in the integrative parts; at a higher level, it means the
restoration of damaged or destroyed parts, as illustrated by the
example of bone regeneration9. Based on the principle that lymph
is the nutritive matter from which organic structures are formed or
reconstituted, in adult as well as in developing organisms, Blumen-
bach distinguishes between two types of forces responsible for the
discharge of fluid in tissues and the assimilation of it by the differ-
ent sorts of parenchymas: on the one hand «the laws of affinity by
which similar parts attract elements of a similar nature»; on the
other hand, «this effort of formation which, by justly applying to
the parts the elementary matter that they need, determines the
precise configuration that molecules of that matter should take»10.
The conjunction of forces gets explained in this way: a fundamen-
tal architectonic disposition determines the organic formation; and
once the structure is set up, nutritive forces of assimilation operate,
which properly depend on complex vital mechanisms. The archi-
tectonic act is what determines the emergence of this or that spe-
cific parenchyma as a sort of constant basis for the nutritive pro-
cesses to take place according to the affinity forces inherent in it11.
In complex and specialized parenchymas, the tendency to self-
reproduction gets so reduced that it may not be capable of regen-
eration anymore. Blumenbach writes:

I believe also, and I conclude from various experiments done on


this subject that in man and the other warm-blooded animals, this
reproductive force is to be found in almost none of the solid and
similar parts that are endowed with contractility and with any other

9 See J.F. BLUMENBACH, Institutions physiologiques, Chez J.T. Raymann, Lyon,


1797, Section 36, De la Nutrition, §459, p. 232.
10 Ivi, §463, p. 234.
11 See Ivi, §461, p. 233: «Quelles que soient ces parties, elles ont pour base

constante, un parenchyme toujours le même. La seule modification qu’elles


éprouvent, leur vient pendant que la nutrition se fait, de la turgescence de leurs
interstices celluleux, que pénètre la lymphe plastique du sang; ou si la nutrition
se fait mal ces mêmes interstices ne recevant pas une assez grande quantité de
cette humeur, s’affaissent sur eux-mêmes, et mentent (sic) l’amaigrissement de
la partie qu’ils concourent à former».
118 François Duchesneau Articles

vital force which is either irritability or sensibility or proper life12.

Once the generative action has produced an organization


that yields specific functional properties, the nisus itself gives way
in the architectonics of the organism to proper modes of conser-
vation and activation of the acquired structures. This state is
correlative to the functional integration of structures in the com-
plex organism. In Ueber den Bildungstrieb, Blumenbach appeals to
this architectonic unity of the integrative structures to account for
the reduction in reproductive capacity of the more specialized
structures:

In man and other warm-blooded animals, the power of repro-


duction is much more limited, not only on account of the great
diversity of materials of which their bodies are composed, but
also of the diversity of the vital forces by which the various
parts of their substance are animated, and of the mutual de-
pendence of these parts on each other. The power of reproduc-
tion is more limited in them than in the polyp 13.

On the contrary, the Bildungstrieb tends to manifest itself


more intensely in relatively undifferentiated organizations, as well
as during the initial phases of embryogenesis from which more
and more complex combinations of cellular, fibrous and other
tissues result.
For direct observation of the effects of this nisus, Blumen-
bach refers to the growth and reproduction of the Conferva fontin-
alis and of the fresh-water polyp (Trembley’s polyp). Theirs are
transparent structures whose growth is sufficiently rapid for one
to follow their morphogenetic phases in continuity. The observed
sequences in the framing-up of such structures do not corrobo-
rate any supposition that pre-existent forms would mechanically
unfold. And so it seems natural to infer that organisms of that
sort form by epigenesis.

12 Ivi, §460, p. 232; cf. also J.F. BLUMENBACH, Manuel d’histoire naturelle, trad. de
l’allemand par Soulange Artaud, Chez Collignon, Metz, An XI – 1803, §19, I,
pp. 35-36.
13 J.F. BLUMENBACH Über den Bildungstrieb, cit., p. 98.
Articles Blumenbach on Teleology 119

When this animal [the polyp] is about to produce its young, a


small tumefaction takes place on a single spot of its body of
plain substance; and from this amorphous, but transparent
swelling, we shall see the cylindrical body of the young polyp
forming, then its tentacles, as if moulded by an invisible hand
from a granular transparent, but formless, jelly. All this is of a
sufficient size to be observed naked-eyed. This fact combined to
the afore-mentioned circumstances does not seem to let any
reason remain for supposing that an organized germ would have
pre-existed therein to develop now14.

This conviction is further supported when one surveys cases


of mutilation of parts with subsequent regeneration of the whole
typical structure. For Blumenbach, these reproductive phenome-
na represent partial rehearsals of the generative process per se. The
best proof of the existence and activity of the nisus is indeed
provided by cases in which segments can reconstitute parts that
have been severed from the individual organism, in which one
can reconstitute the whole polyp by pressing together two halves
of different specimens, for instance, one green and one brown,
and in which the structural reinstatement implies the reusing of
available materials for actualizing a structure that cannot in any
way be reduced to those materials. Blumenbach infers from his
observations that regenerated organs are generally of a lesser size
than the lost originals. He also notes that the regenerative power
diminishes and may even disappear as time goes. These facts
strongly suggest that the implementation of the regenerative force
depends on the organic matter present and available. As a force
inherent in organic matter, the Bildungstrieb is subject to extinction
due to the material conditions in which it operates.
What is at stake here is a means of reconciling the notion of
a final determination towards the framing-up of an integrative
structure with such observed facts as suggest that the epigenetic
process results from the inherent forces of organized matter. In
these conditions, the argument tends to insert so to say the ‘pro-
ject’ or ‘program’ of the organic framework to be achieved within
the material force presumably responsible for the epigenesis. This

14 Ivi, p. 89.
120 François Duchesneau Articles

project or program forms the essential determination of that


force, its ‘form’ in an Aristotelian, or even better, in a Leibnizian,
sense. Hence a fundamental difference with C.F. Wolff’s vis
essentialis: the Bildungstrieb is a force irreducible to the elements
that compose the organism materially, and it exerts a sui generis
organizing power. But it remains the object of a conjecture whose
validity depends only on the observed data for which it provides
a sufficient reason that is as adequate as possible compared to the
available hypotheses:
Because of the weakness affecting all humans, it often happens
that we bring an already-shaped opinion to this type of particular
enquiry, in which the first and least mistake can make us stray
more and more from the way to truth by constant progress on a
divergent line. I felt I would caution myself safely enough against
that kind of error if, having made observations of various type, I
referred them to this issue about generation and carefully exam-
ined whether they all conspired with each other so as to fit one or
the other of the explanatory theories on generation15.

One can easily determine those conspiring phenomena to


which Blumenbach refers in his demonstration. Besides direct
observations of the epigenetic processes in confervas and polyps,
he draws evidence from the production of abnormal membranes,
bones or vessels, which cannot be attributed to pre-existent germs,
but represent an activity organically combining a manifold of
integrative parts. Apart from the Wormian bones that develop
between cranial sutures in hydrocephali, the most telling examples
are afforded by the typical structures arising in parts where they are
not normally to be found (for instance a nail developing on a
finger stump)16 or pseudo-membranes produced in pathological
cases. In abnormal ossifications one cannot suppose that there
would be structural pre-determination for what counts as purely
contingent developments. But one has to admit the equivalent of a
design of functional preservation operating in the accidentally
produced structure: these phenomena require a sui generis formative

15 J.F. BLUMENBACH, De nisu formativo, cit., p. ix.


16 See J.F. BLUMENBACH, Über den Bildungstrieb, cit., p. 99.
Articles Blumenbach on Teleology 121

force that may achieve some functional adjustments17.


Other orders of facts drawn from the formation of hybrids
and monsters are also evoked in support of similar views. Blu-
menbach takes as an example the case of domestic pigs compared
to analogous wild species (boars, in particular). He notes a pro-
pensity for producing anomalies that is more common in con-
texts of domestication. According to the argument for prefor-
mation, it would be impossible to make sense of the fact that
germs have been predetermined to anomalies so as to manifest
themselves under contingent breeding conditions. Besides this
argument, there is a more interesting one according to which the
formative drive and the functional determinants of living organ-
isms would exert themselves differently in changing circumstanc-
es and thus produce more or less adapted structures. This type of
variation, in a non-Darwinian sense, supports the notion of a
power of archetypal production that would be modifiable in its
effects: «How easily the nisus formativus, like other forces of the
animal economy, can be modified following a modification of the
mode of life, is something that does not seem either improbable
or hard to admit»18. Blumenbach tends to suppose that generative
processes tend to be replicated with variations along taxonomic
series, and that the architectonic forces responsible for the
preservation of the type can mutate more or less considerably in
given circumstances. The limit to be found with such mutations
gets proven by the relative impossibility of having fertile hybrids
reproduce, but above all by the requirement for internal harmony
between components for the formation of types:

Upon examination, one easily observes that the organic figure

17 See J.F. BLUMENBACH, De nisu formative, cit., p. xviii: «Ni graviter enim fallor,
in iis calvariis eiusmodi ossicula et maxima et numerosissima deprehendere
mihi videor, quae hydrocephalo interno, vulgatiore sane morbo infantili ac
vulgo creditur, obnoxia fuerunt; quod quidem hydropis genus ut in universum
vehementia valde discrepat, ita maxime mitiores et leviores eius gradus multi-
mode variare videntur : et sæpe quidem sanabiles naturæ medicatricis ope, quæ
huic malo, ossiculorum de quibus iam sermo est, generatione occurrere, et ita
calvariæ distentæ turpitur hiantia interstitia explere et claudere nititur».
18 Ivi, p. xxii.
122 François Duchesneau Articles

imparted to any type of organic bodies, in particular animated


bodies, is the one precisely fitting its mode of life, the functions
dependent upon this figure and the actions that it will undergo19.

Hence there is a relatively strict preservation of the structure


of species, which one must conceive as inherent to the Bild-
ungstrieb as an adequate sufficient reason for type reproduction
through descent. When the mixture of seeds comes from two
individuals belonging to different species, the emergent Bild-
ungstrieb produces a hybrid that possesses all the functional traits
of such an organism. The formative force represents in these
conditions a morphogenetic project that is at once determined
according to natural order and contingent according to the fac-
tors of variation and alteration that prevail in the production of
organic effects. On the other hand, as an explanatory concept,
the Bildungstrieb represents at once the sequential order of epige-
netic developments and the functional order of the organic sys-
tems to be achieved.

2. The Theory of Vital Forces

Blumenbach’s physiological theory presents two significant


features. For one, it follows on Haller’s analytic model and, con-
sequently, several of Blumenbach’s analyses seem to illustrate a
strict correlation between the so-called physiological properties
and the organic parts in which they show up. But on the other
hand, Blumenbach’s physiology puts forward a plurality of specif-
ic vital forces that are directly correlated so as to form a dynamic
system, hierarchically organized and functionally integrated.
Haller conceived of physiology, as an ‘animated anatomy’
(anatome animata), a science of vital motions in their very reality.
The relationship of the integrative organic structures with the
functions exerted seemed to him so complicated to unravel that
he would only determine through observation and experience the
dynamic properties (forces) and functional constants (processes)

19 Ivi, p. xxx.
Articles Blumenbach on Teleology 123

specific to the various structural elements that combine to form


organs and systems. His method of observation and experimenta-
tion would however authorize hypotheses inferred from, and
controlled by, the empirical data. And such hypotheses would rely
on analogical transfers of notions from one field of experience to
another. And thus legitimate physiological theories would arise
from indefinitely perfectible bases.
Among the key hypotheses of Haller’s physiology one finds
the principle that the organism is composed of fibres: «The fibre
is for the physiologist what the line is for the geometer, that from
which all figures arise»20. It is the element of vital organization
that goes into the composition of the various organic structures;
this element varies with the dynamical properties that pertain to it
and determine higher-level functional processes. Haller will not
reduce the functional properties of the elementary fibre to the
sole physical-chemical properties of its compounding material
parts. From the data provided by the analysis of organic struc-
tures and processes, he will rather infer that these units of the
living structure possess specific properties that may account for
more global vital operations, as the properties of an element in a
geometrical figure may account for the functions of its integrative
more complex structures. In the organism it is to this element or
unit that one should refer to establish the relationship between
complex organic functions on the one hand and the ultimately
fine inner structures on the other. For Haller, the levels of inte-
gration in the organism do no strictly correspond anymore to the
mechanical order displayed by nested and juxtaposed micro-
machines; but, taking into account the fibre micro-structures and
their specific functional properties, the physiologist will establish
how membranes that possess such structures are composed and
combined so as to yield the agency and movements that are to be
attributed to organs and organic systems. The exemplary experi-
ments undertaken by Haller at the University of Göttingen from

20 A. VON HALLER, Elementa physiologiæ corporis humani, Sumptibus M.M.


Bousquet & Sociorum (-F. Grasset); Sumptibus Societatis typographicæ,
Lausannæ-Bernæ, 1757-1766 I, p. 2: «Fibra enim physiologo id est, quod linea
geometræ, ex qua nempe figuræ omnes oriuntur».
124 François Duchesneau Articles

1746 on and recorded in the 1752 memoirs De partibus corporis


humani sensilibus et irritabilibus were essentially aimed at a systemati-
zation of vital movements in relation to elementary organic struc-
tures and the dynamic dispositions inherent in their compound-
ing fibres. The irritable fibre is that which, when physically
stimulated, contracts itself with a spontaneous motion, irreducible
to mere elastic contractility. The sensible fibre is that which,
being stimulated, conveys the impression of this stimulation to
the central sense organs and determines such effects in these as
will in turn translate into signs of pain or uneasiness. Haller
experiments systematically in order to establish the typology of
the parts respectively endowed with these distinct dynamic dispo-
sitions and the intensity scales that goes along with them.
If Haller adopted a decentralized view of the organic opera-
tions that rely on the physiological properties of the various
elementary structures, he felt embarrassed in aiming to fix the
theoretical status of these functional properties. In accordance to
the analogy with Newtonian forces of attraction, functional
properties are identified through observable effects that manifest
their presence and specificity. But, while they are presented as
forces inherent (vires insitæ) in those elementary living units which
fibres form, their status as vital forces is left in abeyance: it is
presumed without possible verification that they could be re-
duced to particular internal dispositions equivalent to organic
mechanisms: «A motion cannot take place in the human body
without there being sufficient causes for it in the structure of the
part, and the effect could not be deduced independently of its
cause»21. Could not the phenomena corresponding to irritability
and sensibility boil down to plain results of special mechanical,
that is physical-chemical, material combinations, inner to the
organic units in combination, but properly unattainable? Haller
claims on the one hand that the functional properties of the living
form powers inherent in the elementary structures of organisms;
he supposes on the other hand that an eventual derivation of
these properties from their causes would elicit a «special mecha-

21 A. VON HALLER, Mémoires sur la nature sensible et irritable des parties du corps
animal, M.M. Bousquet (-S. D’Arnay), Lausanne, 1756-1760, I, p. 297.
Articles Blumenbach on Teleology 125

nist hypothesis»22. Confronted with this theoretical dilemma


which he believed he could not solve by resorting to experience,
Haller remained in suspense between positions that will be later
linked to an antinomy between mechanism and vitalism concern-
ing the explanation of vital phenomena.
The functional explanation that relies on identifying physio-
logical properties entails some degree of ambiguity. For explain-
ing the functions of organs and especially of systems that func-
tionally regroup organs boils down to decomposing structures
along morphological elements, identifying the actions and powers
specific to these elementary devices, and presuming that the

22 See F. DUCHESNEAU, La Physiologie des Lumières, M. Nijhoff, La Haye, 1982,


p. 156: «[Chez Haller], la postulation [mécaniste] sous-entend que le
phénomène de l’irritabilité ne serait que la résultante d’un agencement méca-
nique spécial des parties élémentaires. Mais pas n’importe quel agencement
mécanique: le caractère spécial de l’agencement est affirmé, en prenant en
compte que les effets de l’irritabilité sont hétérogènes par rapport aux
phénomènes de l’élasticité et de la contractilité morte. La nature et l’ordre
spécifique des phénomènes vitaux sont sauvegardés dans le cadre d’une
hypothèse mécaniste spéciale». This interpretation has been resumed by other
historians of science. Cf. for instance M.T. MONTI, Congettura ed esperienza nella
fisiologia di Haller. La riforma dell’anatomia animata e il sistema della generazione, Leo
S. Olschki, Firenze, 1990, p. 110: «Natura e ordine sono salvaguardati nel
quadro di un’ipotesi che Duchesneau definisce con molta efficacia
‘meccanicista speciale’. Meccanicista (o meglio: micromeccanicista –
Duchesneau, Ibid., p. 234) – perché la stretta connessione postulata tra
struttura e funzione esclude archei fantasiosi o nature plastiche nella
spiegazione dei dinamismi organici. Speciale, per la riconosciuta specificità che
sussiste fra i diversi tipi di elementi strutturali, fra le vires (da questi innescati) e
fra i rispettivi modelli d’integrazione». See also H. STEINKE, Irritating Experi-
ments. Haller’s Concept and the European Controversy on Irritability and Sensibility 1750-
90, Rodopi, Amsterdam-New York, 2005: «The essential characteristics of
Haller’s concept, however, was neither the maintenance of the correlation
between structure and function nor the division of the body into irritable and
sensible parts but – and this has always been stressed – the postulation of a
strict connection between specific structure and specific function. François
Duchesneau has described this as a ‘special mechanist hypothesis’; mechanist
because of the correlation between structure and function, special because of
both their specificity, and hypothesis because the fibres which should account
for the property were not experimentally accessible, they were only visible in
their more compound formations as nervous and muscular filaments».
126 François Duchesneau Articles

combination of elementary processes along the combination of


microstructures can afford an equivalent for the global processes
which functions properly represent. This part of the conjecture is
meant to project on the morphological elements and their organic
properties, the power to produce combinations of organic parts
at a higher level of complexity; correlatively, the functions that
the machine of nature accomplishes due to its organic composi-
tion are called upon to account for the causal role of the physio-
logical properties in the integrated whole that the organism
forms. Along this line, explaining functions may consist in ap-
pealing to vital forces or principles representing actual physiolog-
ical properties, but surpassing them as they would also potentially
represent emergent functions.
In the analysis he formerly proposed of Blumenbach’s vital-
ism, Thomas S. Hall thus underlined certain features of this
author’s theory of vital forces:

Our prime concern with Blumenbach is centered in his concept


of vitality and of what he terms vital powers. He considered vi-
tality ‘more easily known than defined, [...] its effects [...] ascrib-
able to peculiar powers only’. He considered these powers (vires)
not referable to physical, chemical, or mechanical properties,
however important these properties may be to the economy of
the body. Nor did he acknowledge them to be in any sense
mental, although he saw them as interacting with mental facul-
ties. On the issue of numbers  whether vital powers are one or
many  Blumenbach preferred the latter alternative 23.

If Blumenbach’s views relate to those of Haller, he neverthe-


less tried to work out a new model about the particular economy
of vital forces. The difference between the two methodological
approaches shows up rather clearly. Indeed, vital forces are char-
acterized by the dynamic effects that take place in organic sys-
tems and involve both solids and fluids. Thus vital forces corre-
spond to the organism’s ‘physiological properties’ and stand for
the various presumed agents responsible for the modifications

23T.S. HALL, Ideas of Life and Matter, University of Chicago Press, Chicago,
1969, II, pp. 100-101.
Articles Blumenbach on Teleology 127

and processes occurring in the organism’s structures and especial-


ly microstructures. But, at the same time, the vital forces in their
interrelations are supposed to form an integrative functional
whole. This is what the opening statements of the Institutiones
physiologicae are meant to convey:

In the living human body, whose functions form the object of


physiology, three things are to be taken into account: the solids or
containing parts, the fluids or the humors that are contained in
the former; and what is more important, the vital forces. The vital
forces prepare the solids for receiving the fluids, for expelling
them and for performing several other movements. They are to
be found in the fluids themselves, at least in some of them; and if
they do not form the essences of organized bodies; they are [like]
the soul in them. Though we consider these three things separate-
ly, and though they are really distinct from each other, a living be-
ing conjoins them so intimately, that they cannot be abstracted.
[…] Finally not a single fibre in the living body, however subtle
were it supposed, would be bereft of a vital force24.

Along this approach, Blumenbach provides a theory of the


vital forces as they differ from each other through corresponding
functional acts. This theory outstretches the limits of the strict
structure-function correlations as they appeared in Haller’s physi-
ology. In this theory, properties reducible to inorganic nature are
excluded, even if they are called upon to play a considerable role
in organic functioning – for example elasticity which is character-
ized as «one of the main properties of animal economy»25. But
properties of the soul are similarly excluded, even though they
seem to contribute to the production of certain motions in the
body. The vital forces are those «that exclusively belong to the
organic matter of which we are composed»26. This is a fundamen-
tal restriction and at the same time a theoretical pronouncement,
for forces, be they vital or mechanical or physical-chemical, are
defined as properties of the material sphere alone: there is no way

24 J.F. BLUMENBACH, Institutions physiologiques. Section 1. Du Corps humain


vivant, en général, §§1-2, pp. 1-2
25 Ivi, Section 4. Des Forces vitales en général, §42, p. 24.
26 Ibidem.
128 François Duchesneau Articles

in which we could attribute a non-material status to forces of the


natural order.
From there on, one can proceed to an empirical differentia-
tion, so to say, of specific vital powers. The first is contractility or
cellular force which «residing in the cellular tissue is deployed in its
domain and governs almost all of the body»27. The so-called cellu-
lar tissue is the conjunctive tissue which forms a large set of organ-
ic parenchymas: it was for physiologists in that period, even before
Bichat’s Anatomie générale (1801), the matrix par excellence of all
membrane formations. In normal physiological order, this princi-
ple of contractile force generates phenomena which Stahl used to
refer back to tonicity (motus tonicus): for instance, the phenomena of
tissue absorption, in particular those of lymph absorption as the
nutritive, regenerating and formative element of tissues28. The
second power is Haller’s irritability or muscular force: it properly
belongs to muscle fibres. The third is sensibility or nervous force:
«its effects consist in transmitting to the sensorium the impressions
that affect the organs in which it is present»29; its seat is the nerve
marrow. Blumenbach calls these three powers ‘common forces’
since they are found in all those parts possessing the typical struc-
ture upon which they depend. But this is not the end of the typol-
ogy: under the name ‘proper life’ (vita propria), Blumenbach desig-
nates «the forces in virtue of which only certain parts of our body
fulfill the special functions nature has entrusted them with»30.
‘Proper life’ means here the dynamic raison d’être (principle or force)
that may account for the relationship between a determinate func-
tional operation and a specialized organic structure. Operation in
this case means the integration and correlation of various function-
al acts. The identification of these forces is achieved by differenti-
ating the formal characteristics of the parts concerned in terms of
structures as well as of functions.

27 Ivi, §43, p. 24; see also §50, pp. 26-7: «Déjà nous avons observé que, rési-
dente dans le tissu cellulaire, elle [la contractilité] étend son empire dans
presque tout le corps humain», that is in all parts that originate from the
formation of this tissue: membranes, viscera and even bones.
28 Ivi, §5l, p. 28.
29 Ivi, §45, p. 25.
30 Ivi, §47, p. 25.
Articles Blumenbach on Teleology 129

It is certainly not against sane reason to admit that the parts


whose structure, organization and uses form a class on their
own have been endowed with particular forces. But accurate
observations have shown us, principally in the viscera, forces
that cannot be attributed, in any respect, to the influence of the
common forces31.

Examples given are the movements of the iris, of the Fallo-


pian tubes, of the placenta and of the uterus in delivery, etc., but
also sequences of motions in secretions.
According to Blumenbach, the processes underpinning the
various functions are related, beyond the causal sequences
involved, to a principle capable of integrating the lower level
organic operations for the sake of the effects to be produced;
or, said differently, the causal sequences involved, which may
otherwise be explained in terms of microstructures and organic
motions, form an integrative system of structures and opera-
tions. Hence the hypothesis of a vital force operating in and by
the whole set of its system components and thus providing a
formal expression for the correlation of organic mechanisms.
Thus a particular parenchyma is matched by a specific force
that belongs to a different order from that of the elements out
of which it is made, and different also from that of the vital
operations taking place in these elements. The higher level
specific force may be qualified as emergent.
The fifth and last type of force is the Bildungstrieb itself:

[This formative force] is the efficient cause of every conserva-


tive and reproductive act. It is by means of it that generative and
nutritive matters are introduced into appropriate reservoirs,
where they are adequately worked upon, undergo the necessary
transformations, and are finally changed into parts that are ca-
pable of contractility, irritability, sensibility, or proper life 32.

As a power of framing up the available organic matters, the


formative force emerges from appropriate combinations of these

31 Ibidem.
32 Ivi, §48, p. 26.
130 François Duchesneau Articles

matters and translates into organogenesis the specific order of a


given type of organism. As an organizing force, it causes the
subsequent emergence of more specialized vital forces.
In his polemics with Caspar Friedrich Wolff, Blumenbach
took advantage of his analysis of tissue nutrition to cope with the
question of the plurality of forces in the model he proposed. Are
we dealing with so many intrinsically distinct vital forces or with
the several modalities of a single force? His preference went to a
plurality or even a multiplicity of dynamic principles. Against the
speculative tendency to artificially simplify, he seemed to favour
the conception of units of vital force spread throughout organic
matter, but capable of combining their respective actions to
achieve integration and functionality. In this instance, he was
content with presenting these forces as means for the true ‘vis
essentialis’, which the Bildungstrieb is33. What lies beyond any possi-
ble explanation is the nature of these various forces as causes of
the laws that express their efficiency, in particular when this
efficiency results in programs of functional organization. These
programs involve the various parts and operations of the living in
mutual relation, from their initial formation to the achievement of
their structures and the emergence of their specialized functions.
The ‘force of nutrition’ (Nutritionskraft) signifies according to
Blumenbach a form of physical-chemical processes, specific to
organized matter; these processes provide necessary, but not
sufficient, conditions for the formation and preservation of the

33 P. McLAUGHLIN, Blumenbach und der Bildungstrieb, «Medizinhistorisches


Journal», XVII, 1982, p. 364-366, reminds us that Wolff presented his vis
essentialis as a force belonging to the general mechanical order of Nature and
capable of forming organic structures from inorganic matters. The action of
this force might be ultimately analyzed into physical-chemical processes due to
specific affinities. On the contrary, the Bildungstrieb would appear as an exclu-
sively vital force that would deploy itself only in a matter already endowed with
organic properties. If this could be the reductionist ontological ground of
Wolff’s vis essentialis, assimilated to a special mechanical property of matter in
general, it is not sure, in my opinion, that the analysis of the sequences of
effects arising from the action of this force would not characterize a specific
order of vital functions. Thus the non-reductionist reinterpretation Blumen-
bach gives of this force in his memoir Ueber die Nutritionskraft could gain some
additional credit.
Articles Blumenbach on Teleology 131

organism. The physiologist must in addition suppose that these


processes achieve a dynamic set of dispositions combining the
various modes of organic assimilation and secretion. Such opera-
tions take place as if these functional effects would depend upon
a program-oriented or program-directed set of organic motions.
This goal-directedness is revealed by the organ-generating power
of the Bildungstrieb, which in turn is the source and origin of those
vital forces fulfilling specific organic functions. In my opinion,
Blumenbach’s model associates, as Kant seems to have rightly
understood, «the physical-mechanical principle with the purely
teleological mode of explaining organized Nature»34.
The causal explanation of physiological phenomena proceeds
within a framework of ‘special physics’, by presuming that a kind
of ‘regulation’ rules over organic dispositions and processes as
over a set of elementary causal sequences. This ‘regulation’ entails
an integrative arrangement of parts as well as operations that are
appropriate for eliciting and maintaining the functions of the
whole, starting from the primordial functions of generation and
conservation of the vital unit(s). As suggested in the Commentatio
de vi vitali sanguinis (1788), one should not postulate any vital force
where one cannot determine the functionality of the effect rela-
tive to the production and preservation of an operational organic
order. The effects resulting from vital forces (contractility, irrita-
bility, sensibility, proper life, nisus formativus) precisely reveal such
functional properties.

What end the diverse orders of vital forces in the solid parts of
our body fulfill, a close examination shows it clearly. Surely con-
tractility enables the cellular tissue to push the fluids inherent in
its compounding cells, etc.; irritability is of use to muscular fibre
in order for muscles to be able to accomplish by far a majority
of movements, especially voluntary ones; sensibility of the nerve
marrow helps convey to the sensorium the impressions from
stimuli; proper life enables the uterus, the iris and the other
parts of that kind to exert special functions. But blood as such
does not seem to need [vital] force, and in general none of the
animal’s humors has to change into solid parts, except the nisus

34Letter of Kant to J.F. Blumenbach, 5 August 1790, in: Kants Werke. Akademie
Textausgabe, Walter de Gruyter, Berlin, 1968, XI, p. 185.
132 François Duchesneau Articles

formativus: and [according to our demonstration] this applies only


to the generative fluid and plastic lymph35.

The formal feature of functionality shows up in addition


when one compares the specific effects of vital processes with the
corresponding pathological phenomena: the forces that emerge
from the various systems are thus highlighted by the alterations
affecting the corresponding organic operations.
This being so, any analysis of vital functions can only be rela-
tive, for the order of phenomena results from a manifold of
structural and functional conditions nested in one another at
various levels of integration. As stated by Blumenbach:

From the agreement that prevails among the solids, fluids and vi-
tal forces, from the sympathy that links the numerous divisions of
which we are composed, finally, from the close union between
body and soul, life and health result, which are two attributes
whose degrees are multiplied and diversified to infinity36.

However, the state of balance achieved by this harmonious


convergence among material systems and intervening forces is
precisely determined according to the type of organism. Forces
intervene in physiological processes by typical sequences of
actions and reactions: these sequences express the effects of the
underlying forces; and these effects, though multiple and diversi-
fied, are shown to be constant and regular once they are related

35 J.F. BLUMENBACH, Commentatio de vi vitali sanguinis, J.C. Dieterich, Gottingæ,


1788, p. 7: «Cui porro bono virium vitalium diversi ordines in solidis corporis
nostri partibus inserviant, attente rem perpendenti facile patet. Cellulosæ sc.
telæ in universum contractilitas ad pellenda fluida cellulis eius inhærentia etc.;
muscularis fibræ irritabilitas ut longe plurimi motus, maxime voluntarii, muscu-
lorum ope peragi possint; nervosæ medulæ sensilitas ad stimulorum impressiones
sensorio communicandas; utero, iridi, aliisque id genus partibus vita propria ad
peculiares functiones subeundas. Sanguis vero qua talis nulli vi egere videtur, et
in universum nullus humor animalis, nisi nisu formativo in solida vertendus:
quod unice de genitali latice æque plastica lympha valere infra dicetur».
36 J.F. BLUMENBACH, Institutions physiologiques, Section Cinquième. De la Santé

et de la Nature humaine, §56, p. 34.


Articles Blumenbach on Teleology 133

to the functions to be performed37.


What is epistemologically remarkable with this type of model
is the status of vital principles qua sufficient reasons for organic
operations beyond the level of structures and motions. Every-
thing takes place as if the functional order of effects was project-
ed in the form of sui generis principles beyond the systems and
activities the organism consists in. Everything takes place as if the
organic order over-determined the elements, structures, and
mechanical, physical and chemical factors involved. Blumenbach
seems to aim at projecting a functional determining reason, or
more exactly, a series and even a hierarchy of functional laws
over the physical-chemical means of organic activity. This ap-
proach boils down to ascertaining the existence of an architecton-
ic principle, the Bildungstrieb, and of principles deriving from it,
that is vital powers common or specific to elementary as well as
compounded organic structures.
In this instance, some of the ‘axioms’ of the Handbuch der
Naturgeschichte are meaningful. This is particularly the case with
Axiom I: the organism’s vital forces are intimately dependent
upon the organization of the parts and of the whole where they
are deployed38. And it is also with Axiom IV: since «the idea of an
organized body which necessarily implies the idea of an end, of a
goal, absolutely defeats all purely mechanical explanations of the
successive formation of organized bodies»39; the Bildungstrieb
stands as a specific force responsible for the structural and func-
tional integration of the living body in accordance with the idea
one legitimately conceives of it, while serving at the same time as
a natural agent of organic formation. And this justifies stating:
«[The name Bildungstrieb] should only designate this particular

37 See Ivi, §53, p. 31: «Les solides, les fluides et les forces vitales agissent et
réagissent perpétuellement les uns contre les autres dans le corps humain vivant.
Les fluides agissent en stimulant les solides, ceux-ci, en vertu de la force vitale
dont ils sont doués, ressentent cette action, et réagissent contre elle. Ces
alternatives d’actions et de réactions, ces mouvements opposés se balancent
dans un homme sain, et se maintiennent dans un état d’équilibre précis».
38 See J.F. BLUMENBACH, Manuel d’histoire naturelle, §6, p. 15. Concerning

reproduction see §5 et §20.


39 Ivi, §8, note, p. 20.
134 François Duchesneau Articles

force that combines in itself the mechanical principle with this


other one which is subject to modifications conformably to an
end»40. As for Axiom V, it states that, while achieving this archi-
tectonic project, the Bildungstrieb entails the various functional
powers that will be later actualized in and by organ-dependent
processes. To sum up, the concepts of the Bildungstrieb and of the
vital forces deriving from the formative nisus would account for
the organic phenomena by subsuming under an end of vital
organization those mechanical, physical and chemical processes
that the living body deploys in and by itself in reaction to various
sets of external conditions. Consequently, appropriately framed
theoretical concepts would direct our understanding of physio-
logical phenomena by relating the causal sequences displayed in
the organism to the integrative processes involved in the living
body’s formation and operations.

3. Conclusion

With a few total or partial eclipses, this model, duly trans-


posed, prevailed in most of German physiology throughout the
19th century, and probably much beyond. It is especially noticea-
ble in embryology conceived as Entwicklungsgeschichte and in the
framing-up of the cell theory. The research programs that fol-
lowed upon its inspiration would conjoin reductionist strategies
in analyzing and explaining vital processes with presuppositions
about a specific order or organization ruling over the combina-
tion of processes. The core version of this synthetic approach
consisted in projecting as theoretical entities sui generis forces that
would achieve the integration of a manifold of physical and
chemical ‘analytic processes’.
I entertain no doubt that in its time this so-called vitalist
model did fulfill the objective of eliciting laws for the functioning

40 Ivi, §9 (2nd remark), p. 23. See J.F. BLUMENBACH, Handbuch der Naturgeschich-
te. bey Johann Christian Dieterich, Göttingen, 1799, p. 18: «[...] sondern bloß
eine besondre (das Mechanische mit dem zweckmässig Modificirbaren in sich
vereinende) Kraft unterscheidend bezeichnen soll».
Articles Blumenbach on Teleology 135

of organisms. This has been for instance the case in embryology


with Karl Ernst von Baer and Robert Remak, in physiology with
Johannes Müller, in pathology with Virchow and his followers,
but, also beyond, in the various spheres of biology still in the
becoming wherein many direct or indirect followers of Blumen-
bach left their mark.
An interesting case of positivist and deterministic adoption
of a similar framework is afforded by Claude Bernard’s complex
explanatory model. In a former paper devoted to the impact of
the cell theory on Bernard’s general physiology41, I mentioned
that, mutatis mutandis, Bernard projected a methodological scheme
according to which general physiology would be analytic only
when investigating determinisms derived from the activity of cells
duly formed. Bernard’s reductionism would take as a requisite
that an elementary organism has been previously formed whose
generation could not be subject to experimental analysis stricto
sensu. Forming a kind of epistemological restriction, this adventi-
tious hypothesis denoted a properly synthetic dimension of the
program of general physiology. William Coleman considering the
cognitive basis of Bernard’s physiology once wrote: «The crucial
matter was to accept the purposiveness of life as a principle of
analysis, but also to insist upon the legitimacy and necessity of
approaching living organism in a deterministic manner»42. This
stated ambiguity may have been on the whole relatively fruitful as
a methodological stance in approaching the study of such highly
complex phenomena as those of vital organization, for, notwith-
standing the teleological concepts involved, it favoured in parallel
the expansion of deterministic analyses of the powers and prop-
erties of organic bodies.

41 F. DUCHESNEAU, Claude Bernard et le programme de la physiologie générale. in: C.


Blanckaert et al. (Eds.), Le Muséum au premier siècle de son histoire, Éditions du
Muséum National d’Histoire Naturelle, Paris, 1997, pp. 425-445.
42 W. COLEMAN, The Cognitive Basis of the Discipline: Claude Bernard on Physiology,

«Isis», LXXVI, 1985, pp. 49-70.


TELEOLOGY BEYOND REGRETS:
ON THE ROLE OF SCHELLING’S ORGANICISM IN
TREVIRANUS’ BIOLOGY

by Andrea Gambarotto

Abstract. In his seminal monograph on teleology and mechanics in nine-


teenth German biology Timothy Lenoir considers his study of the “Kantian”
teleomechanistic tradition as an answer to those who wrongly believe that early
nineteenth-century German biology was dominated by Schelling’s Natur-
philosophie. My goal is to argue that this is an arbitrary assumption based on
a historiographical bias and that Schelling’s organicism played a pivotal role
in the formulation of a conceptual framework aimed at accounting for biologi-
cal organization. The formalization of biology as an autonomous science at
the beginning of the nineteenth century implied in fact the shift from a regula-
tive to a constitutive understanding of teleology, a shift most strongly endorsed
in Schelling’s Naturphilosophie. I first take into account two aspects that
Treviranus draws directly from Schelling: the relationship between mechanism
and teleology and the continuity between nature and spirit. I then show how
Treviranus reinterprets the Schellingian framework with a peculiar emphasis
on ecology, stressing the important interaction between organisms and envi-
ronment. On this basis, I suggest that he was the first naturalist in the
German speaking world to sketch the outline of a theory concerned with the
historical transformation of living forms.

0. Introduction*

The title of this paper refers to an old controversy in the his-


tory and philosophy of biology that originated in the scholarly
work of Timothy Lenoir. In a paper entitled «Teleology without

*Many thanks to Giovanni Menegalle and Charles Wolfe for reviewing previ-
ous drafts of this paper.
Verifiche XLIII (1-4), 2014, pp. 137-153.
138 Andrea Gambarotto Articles

regrets»1 Lenoir discusses the main features of the account he had


elaborated in a series of papers on the same topic2, and which
would soon converge in his seminal monograph on teleology and
mechanics in nineteenth-century German biology3. According to
the standard account, he maintains, the real beginnings of scien-
tific biology are best exemplified by the efforts of the ʻ1847
groupʼ (Ludwig, Du Bois Reymond, Helmholtz and Brücke), who
threw off the yoke of «vitalistic explanation» and swore allegiance
to the cause of «mechanistic reductionism» (neither of these
terms is given a clear definition). Vitalism and teleology had
thereby been cast aside and the reign of mechanistic biology
inaugurated4. Although, according to Lenoir, this account implies
that the only consistent foundations of biology are those supplied
by mechanistic reductionism, he argues that a coherent and well-
developed research program guided the development of the life
sciences in Germany from the 1790 through the mid-1850s. He
defines this program as «teleomechanism» and analyzes it in terms
of three different phases: «vital materialism» (Kant, Blumenbach,
Reil, Kielmeyer), «developmental morphology» (Meckel, Döllinger,
von Baer, Müller), and «functional morphology» (Schwann, Liebig,
Bergmann, Leuckart).
In his attempt to analyze the first phase Lenoir employs the
category of ʻGöttingen Schoolʼ. This category has the merit of
stressing the existence of a unitary center characterized by intense
institutional and intellectual relations among nearly three genera-
tions of physicians and naturalists. According to Lenoir, the
distinctive approach practiced at Göttingen derived from ideas

1 T. LENOIR, Teleology without Regrets. The Transformation of Physiology in Germany:


1790-1847, «Studies in History and Philosophy of Science Part A», 12 (4),
1981, pp. 293-354.
2 T. LENOIR, Generational Factors at the Origin of “Romantische Naturphilosophie”,

«Journal of the History of Biology», 11 (1), 1978, pp. 57-100; ID., Kant, Blumen-
bach and the Vital-Materialism in German Biology, «Isis», 70, 1980, pp. 77-108; ID.,
The Göttingen School and the Development of Transcendental Naturphilosophie in the
Romantic Era, «Studies in the History of Biology», 5, 1981, pp. 111-205.
3 T. LENOIR, The Strategy of Life: Teleology and Mechanics in Eighteenth Century

German Biology, Riedel, Dordrecht 1982.


4 T. LENOIR, Teleology without Regrets, cit., pp. 293-294.
Articles Teleology beyond Regrets 139

fashioned by Blumenbach, who synthesized some of the best


elements of Enlightenment life sciences, especially Buffon, Lin-
naeus and Haller, in terms of a view of biological organization he
found in the writings of Kant. Blumenbach graduated at Göttin-
gen in 1776 and was appointed professor in 1778. His reputation
was considerably enhanced by the publication of his Institutiones
Physiologicae, a condensed, well-arranged view of the animal func-
tions. His physiological theories established the foundations of
the Göttingen School and were developed by his most distin-
guished students: Alexander von Humboldt, Johan Christian Reil,
Carl Friedrich Kielmeyer, Heinrich Friedrich Link, Gottfried
Reinhold Treviranus5.
William Bechtel has pointed out that the aim of Lenoir’s re-
construction was to identify a tradition in nineteenth-century
German biology different both from vitalistic Naturphilosophie and
reductionist materialism6. Released from its entanglement with
vitalism, teleology could finally be considered in naturalized terms
(i.e. without regrets), as a specific characteristic of organic pro-
cesses marking the irreducibility of biological phenomena to
physics and chemistry. Lenoir sees his study of the ʻKantianʼ
teleomechanistic tradition as an answer to those who wrongly
believe that early nineteenth-century German biology was domi-
nated by Schelling’s Naturphilosophie and by its ʻvitalisticʼ concep-
tion of teleology. This reconstruction has been harshly criticized
by Kenneth Caneva in a review entitled, ironically enough, «Tele-
ology with Regrets», where Lenoir is accused of «many serious
mistakes in historical analysis», «errors, misinterpretations, incon-
sistencies, unsupported claims and plain unclear writing»7. It is in
fact odd to maintain that the «vital-materialism» developed at
Göttingen rejected the vitalistic notion of purposive activity,
given that Kant’s conception of teleology was intimately tied to a
notion of purposiveness, as was Blumenbach’s Bildungstrieb8. More-
5 T. LENOIR, The Göttingen School, cit., pp. 115-119.
6 W. BECHTEL, Teleomechanism and the Strategy of Life, «Nature and System», 5,
1983, 181-187.
7 K. CANEVA, Teleology with Regrets, «Annals of Science», 47, 1990, pp. 291-300,

p. 300.
8 See the paper by François Duchesneau in this volume.
140 Andrea Gambarotto Articles

over, many of Lenoir’s ʻteleomechanistsʼ broke with a severe


Kantian notion of teleology as a merely regulative concept of the
understanding and instead conjured up a variety of vital forces
more or less actively constitutive of the individual organism9. Stress-
ing the heuristic value of teleological thinking in biology only
inasmuch as it can be reduced to a mechanistic framework of
explanation, Lenoir acknowledges a role for teleology but indeed
he does so ʻwith regretsʼ.
I intend to argue that we are in need of a new general ac-
count going beyond those regrets. In fact, I will argue that that
the formalization of biology as an autonomous science at the
beginning of the nineteenth century implied the shift from a
regulative to a constitutive understanding of teleology – a shift
most strongly endorsed in Schelling’s Naturphilosophie. Biology as a
science became possible only once organization was considered
as a constitutive character of living bodies which, as such, re-
quires scientific explanation.
The term ʻbiologyʼ has traditionally been traced back to La-
marck and Treviranus, who first used it in significant fashion in
1802 in the Recherches sur l’organisation des corps vivants and in the
Biologie, oder Philosophie der lebenden Natur für Naturforscher und Ärzte.
In fact, other authors had already used the term en passant slightly
earlier, such as Georg August Roose in the Lehre von der Grundzüge
der Lebenskraft (1797) and Karl Friedrich Burdach in his Propädeu-
tik zum Studium der gesamnten Heilkunst (1800). As scholars have
recently stressed, the word itself was used in the sense of
ʻbiographyʼ even earlier, so that we may have to move the date of
the first use of the term another thirty years. Michael Christoph
Hanov, a minor disciple of Christian Wolff, published from 1762
to 1768 a four-volume Latin compendium entitled Philosophia
naturalis sive physica dogmatica, whose third volume (1766) bore the
subtitle: Geology, Biology, General Phytology and Dendrology, or the

9 Cf. R.J. RICHARDS, Kant and Blumenbach on the Bildungstrieb: A Historical Misun-
derstanding, «Studies in History and Philosophy of Biological and Biomedical
Sciences», 31 (1), 2000, pp. 11-32; J. ZAMMITO, The Lenoir Thesis revisited: Blu-
menbach and Kant, «Studies in History and Philosophy of Biological and Bio-
medical Sciences», 43 (1), 2012, pp. 120-132.
Articles Teleology beyond Regrets 141

Science of the Earth, of Living Things and of Vegetating Things in General,


as well as of Trees. However, if one discounts the running heads,
Hanov does not use the word ʻbiologyʼ in the text of the volume
itself. There are even more minor Wolffians that could be taken
into account, but «none of this really affects the more important
question of the mechanisms of the historical development and
institutionalization of the life sciences in the nineteenth century»10.
The grand baptism of the concept is still to be located in the
writings by Treviranus and Lamarck. In this respect, if a lot has
been said on Lamarck11, almost no word has been uttered about
the other pioneering endeavor, in which the idea of a unified
science of life is endorsed with the strongest arguments. The
reason for this silence has two explanations. The first is almost
certainly the magnitude of the opus, since a nine-book treatise
divided into six volumes (each around five-hundred pages) poses
an obvious challenge to scholarly work. Secondly, the over three-
thousand pages that compose this work are filled with references
to countless eighteenth-century scientific literature and, even
more importantly, with a language that can be properly under-
stood only by being well acquainted with the jargon of eight-
eenth-century German life sciences.
Before I begin my analysis, it is useful to have in mind an
overall sketch of the general division of the work. As mentioned,
the Biologie is composed by nine books, divided into six volumes;
(1) The first volume includes a long introduction, where Trevi-
ranus defines the fundamental concepts and the theoretical
framework of biology as a new scientific field, and the first book
of what he refers to as «history of physical life» (Geschichte des
physischen Leben), dedicated to the general «classification of living
organisms»; (2) The second volume consists of the second book
on the «organization of living nature», where Treviranus provides
a detailed account of the distribution of living bodies on the
10 P. MCLAUGHLIN, Naming Biology, «Journal of the History of Biology», 35 (1),
pp. 1-4, p. 4.
11
G. BARSANTI, Dalla storia naturale alla storia della natura. Saggio su Lamarck,
Feltrinelli, Milano 1979; P. CORSI, The Age of Lamarck. Evolutionay Theories in
France: 1790-1830, University of California Press, Berkeley 1988; P. CORSI, J.
GAYON, J.G. GOHAU, S. TIRARD, Lamarck, philosophe de la nature, Puf, Paris 2006.
142 Andrea Gambarotto Articles

different areas of the earth, depending on different environmen-


tal conditions; (3) The third volume contains the third and the
fourth book of the history of physical life: the former is con-
cerned with the revolutions that occurred to living nature over
time, while the latter is dedicated to the exposition of Treviranus’
theory on «generation, growth and decrease of living bodies»; (4)
The fourth volume is occupied by the fifth book and is con-
cerned with the formulation of a general theory of nourishment;
(5) The fifth is concerned with physiological issues and entails
three books (sixth, seventh, eighth) dedicated to «warmth, light,
and electricity of living bodies», to the «automatic movement of
living bodies», and the «functioning of the nervous system»
respectively; (6) The sixth is dedicated to the «connection of the
physical with the intellectual world» and provides an outline of
brain physiology in the animal kingdom.
The overall work provides perhaps the best example of the
sedimentation of the conceptual framework elaborated at Göt-
tingen and developed by Naturphilosophie. In this respect, to char-
acterize the Biologie as a ʻground-breakingʼ work would probably
be an overstatement. Nevertheless, despite its compilatory nature,
this massive collection of materials is the final result of a concep-
tual course concerned with the endeavor of providing an ade-
quate definition (and a corresponding explanatory framework) to
the way living nature is capable of organizing itself. Roughly
speaking, this course can be said to begin with the Haller-Wolff
debate12 and culminates with Schelling’s idea of nature as a «uni-
versal organism», i.e. as a dynamical system capable of organizing
and regulating itself. Despite its being defined as «dynamic evolu-
tion», this self-organization is understood by Schelling as a non-
temporal process. The production of living forms is the result of
the eternal tendency of nature to give exposition to the absolute, a
task that, according to Schelling, can never be completely fulfilled.
Among German naturalists, Treviranus was the first author to
place this process on a temporal plateau: this happens in the third
book (entailed in the first part of the third volume), which, for this

12Cf. S.A. ROE, Matter, Life, Generation. Eighteenth-Century Embryology and the
Haller-Wolff Debate, Cambridge University Press, Cambridge 1981.
Articles Teleology beyond Regrets 143

reason, is the most relevant of all. As with Lamarck in France,


Treviranus was the first author to endorse at once the scientific
autonomy of biology and a consistent theory of tranformism.
I will provide evidence for these claims in two steps: (1) I will
take into account two aspects that Treviranus draws directly from
Schelling’s Naturphilosophie: the relationship between mechanism
and teleology and the continuity between nature and spirit; (2) I
will then show how Treviranus reinterprets the Schellingian
framework with a peculiar emphasis on ecology, stressing the
important interaction between organisms and environment. On
this basis, I will suggest that he was the first naturalist in the
German-speaking world to sketch the outline of a theory con-
cerned with the historical transformation of living forms.

1. Naturphilosophie behind Biology

1.1. Mechanism and Teleology

The first element I will take into account is the relationship


between mechanism and teleology. This is a pivotal theme in the
natural philosophy of German Idealism, which originated in
Kant’s Critique of the Power of Judgment. In the second part of this
work, dedicated to the «teleological judgment», Kant addresses
the problem of whether the specific form of a living being can
be judged as being organized according to specific purposes. The
form of a bird, especially its bone structure and the position of
its wings, suggests a positive answer to the question, as their
intrinsic purpose seems to be that of flight. To Kant, however,
this is the same as to conceive nature in technical terms, i.e. as if
it were the product of a maker, because «nature, considered as a
mere mechanism, could have formed itself in a thousand differ-
ent ways without hitting precisely upon the unity in accordance to
such a rule»13. This assumption is at the basis of the so-called
antinomy of teleological judgment: on the one hand, in fact, «all

13 I. KANT, Kritik der Urteilskraft, Akademie Ausgabe, 5, 360; Critique of the Power
of Judgmrnt, Cambridge University Press, Cambridge 2000, p. 234.
144 Andrea Gambarotto Articles

generation of material things is possible in accordance with


merely mechanical laws», while on the other hand «some genera-
tion of such things is not possible in accordance with merely
mechanical laws»14. The Kantian solution to this dilemma is the
introduction of the distinction between determinant and reflective
judgment. The former refers to a constitutive property of the
object called into question, the latter to the way in which our
cognitive faculty makes sense of things. According to Kant we
must consider living organisms as if they were the product of
intentionally acting causes, while dealing with them in a mecha-
nistic framework of explanation15.
In his early philosophy of nature, Schelling was the first to
challenge this view. In the Erster Entwurf eines System der Natur-
philosophie (1799), mechanism, chemical affinity and teleology are
considered as different «powers» (Potenzen), which characterize
different levels of the natural system. At the lowest levels, ele-
mentary compounds are extrinsic from one another and interact
only through mechanical relations; at a higher level, magnetism
and chemical affinity testifies to the existence of more intrinsic
interactions, in which the relation between the terms in play
determines their proper character; finally, the realm of living
organisms is altogether holistic, as the whole thoroughly deter-
mines the structure and the function of the single parts. This
framework received a most schematic formulation in the section
Objectivität of Hegel’s Science of Logic, which is explicitly divided
into three parts entitled «mechanism», «chemism» and «teleology».
The idea behind this schema is that the teleological features
manifested by living organisms is not a mere imputation of our
faculty of judgment, as claimed by Kant, but rather a constitutive
property of their structure.
In the fourth volume of the Biologie Treviranus maintains that
in order to satisfactorily account for the functions of living bod-
ies it is necessary to adopt a new point of view. This point of
view implies a shift from a mechanical to a teleological approach.

Ivi, p. 387; p. 259.


14

Cf. I. GOY & E. WATKINS (eds.), Kant’s Theory of Biology, DeGruyter, Berlin-
15

New York 2014.


Articles Teleology beyond Regrets 145

That is, one has to understand the functioning of living organ-


isms as acting in accordance to the specific purpose of maintain-
ing the general structure and overall organization of the living
body. Treviranus maintains that teleology must be envisaged as
the truth of mechanism, i.e. as an explanatory principle necessary
to explain organic nature, a realm characterized by higher organi-
zational complexity. If the principle of mechanism can be em-
ployed to account for the phenomena of physics, biological
entities must in fact be considered according to the principle of
purposiveness, i.e. their functions must be understood as serving
specific purposes. It is fruitless to look for an explanation of life
if we do not assume that its emergence «is grounded on a princi-
ple to which must be ascribed a certain degree of independence
from external influences, of self-determination to effectiveness,
analogue to spontaneity». The vital principle (Lebensprinzip) of
every organism is grounded in a common fundamental force
(Grundkraft) through which «living nature displays phenomena
whose cause is higher than mere mechanical or chemical pow-
ers»16. Chemical principles can explain the elements of which
living bodies are composed. One can also investigate «all the
traces of electricity, magnetism and all further physical forces in
the living body and pursue them as far as possible» but «the actual
secret of living nature will not be disclosed»17. The realm of living
organisms is different from those of mechanism and chemical
affinities and requires a reference to teleological principles.
This notion of teleology as an inherent property of biologi-
cal beings is the first element linking the conceptual framework
of Treviranus’ Biologie to that which was first elaborated by the
philosophy of nature of German Idealism in contrast to Kant. It
is hard to see how the striking similarity between the two could
be explained otherwise than through their belonging to the same
discursive configuration. The way in which Treviranus articulates
the relationship between nature and mind provides further argu-
ments in favor of this thesis.

16 Ivi, p. 629.
17 Ivi, p. 631.
146 Andrea Gambarotto Articles

1.2. Nature and Spirit

The purposive characteristics displayed by living organisms


are for Schelling a first manifestation of the intrinsic spiritual
character of nature. This character is completely concealed in the
phenomenon of mechanism and starts manifesting itself in
chemical affinity and magnetism, as the relation between the
terms in play becomes more internal and less extrinsic. It is how-
ever at the biological level that the spirit within nature is fully
manifested. A first mark of this spirituality is the fact that living
organisms display a form of independence and spontaneity,
which is absent in mechanical and chemical phenomena. This
spontaneity is testified by animal instincts and completely realized
in human consciousness. A further effect is for Schelling the
«technical drive» (Kunsttrieb), which is «only the final work of the
same force that produced the organism itself»18.
Treviranus formulates a very similar argument, claiming that
organisms display characteristics which have a very determinate
purpose: the instincts or natural drives (Naturtriebe). These can
relate either to the individual or to the genus. In the former case
they include the drives of self-preservation and self-defense, in the
latter the drive to reproduction. They all «have in general the char-
acter of purposiveness (Zweckmässigkeit)»19. The waking of instincts
is a result of «the continuing and partially modified activity of the
original formative drive, the only one among the vital forces (Le-
benskräfte) which, like the instincts, displays purposiveness and an
appearance of spontaneity in its effects»20. As with Schelling,
Treviranus considers the ʻmindʼ to be an internal development of
nature which manifests its intrinsic teleological features.
The discussion of instincts leads in fact towards a «domain
full of obscurity»: the intellectual faculties of living organisms.
Treviranus maintains that, throughout history, scholars expressed

18 F.W.J. SCHELLING, Erster Entwurf eines Systems der Naturphilosophie, From-


mann-Holzboog, Stuttgart 2001, p. 202.
19 G.R. TREVIRANUS, Biologie, oder Philosophie der lebenden Natur für Naturforscher

und Ärzte, V, Röwer, Göttingen 1818, p. 430.


20 Ivi, p. 443.
Articles Teleology beyond Regrets 147

two opposed views on this issue: either spirit and matter were
considered completely different in nature or as related to one
another. By contrast, he claims that life «lies in a principle, whose
essence is self-activity». The use of this notion is very innovative
in a biological context, as much as it is frequent in the vocabulary
of German Idealism. This self-activity «expresses itself as forma-
tive drive and is merely immanent. It persists also in the formed
organism and expresses itself through further formation and
preservation»21. Autonomy is the fundamental character of animal
life. An organism that displays this autonomy behaves «with the
appearance of conscience and freedom, but nevertheless uncon-
sciously and according to necessary laws»22. Generally speaking,
memory is the most widely shared intellectual capacity in the
animal kingdom. Treviranus considers bees that every year return
to the place where they had been fed the previous summer. This
ability «is not possible without imagination (Einbildungskraft),
which must thus befit animals». If the use of this term were not
sufficient, the footnote to this page explicitly refers to the corre-
spondence between Jacobi and Fichte.
Certain «technical drives» (Kunsttriebe) must also be ascribed
to animals, because otherwise it would be impossible to explain
the construction of artifacts in the animal kingdom, such as the
building of a nest. This is the same notion employed by Schelling
in the Erster Entwurf. Although, in fact, humans are different from
animals, they are not so different that similarity is eliminated. This
is because «the degrees (Stufen) that humans go through from
their origin to their complete formation» can be compared with
«the degrees of development of the animal kingdom from infu-
soria to human beings»23. This is a reformulation of the recapitu-
lation theory sketched by Kielmeyer and developed by Oken
(later known as the Meckel-Serres law). According to Treviranus,
moreover,

the same force that forms living bodies from formless matter,

21 Ivi, p. 5.
22 Ivi, p. 6.
23 Ivi, p. 24.
148 Andrea Gambarotto Articles

works in them as a conservative and healing force of nature af-


ter their formation, expresses itself as instinct and on the spir-
itual side, as imagination, it is the producer of ideas24.

This idea of continuity between nature and spirit is strikingly


similar to that endorsed by Schelling, and later by Hegel25.
The textual evidence I have provided so far should be suffi-
cient to sustain the thesis that the framework elaborated by
Naturphilosophie played an important role in laying the foundations
for the emerging biological science. This is especially true of
Schelling’s organicist views, according to which nature and mind
are conceived of as different developmental levels within a single
natural system. The various degrees connecting these two include
mechanism, magnetism, chemism and teleology. This framework
implied a crucial shift from the Kantian understanding of teleol-
ogy as a mere regulative principle to a conception of teleology as
a constitutive property of living beings. In the following section I
explain how Treviranus adopted this schema and extended it in
relation to the empirical research of his time, thereby elaborating
a consistent theory of transformism.

2. Biology beyond Naturphilosophie

Despite these convergences there is one pivotal difference


between Schelling and Treviranus. For the former the succession
of different degrees of organization is conceived of as merely
logical, i.e. it describes the way nature is synchronically struc-
tured and accounts for its internal articulation. This articulation
itself never changes. Schelling sees nature as composed by a
series of degrees which progressively manifest its intrinsic
spiritual character. Treviranus adopts this framework but inter-
prets it in diachronic terms.
The emphasis on the geographical distribution of organic
life was fundamental for this reinterpretation. In fact, for Trevi-

24 Ivi, p. 28.
25 See the paper by Luca Illetterati in this volume.
Articles Teleology beyond Regrets 149

ranus the distribution of animals does not take place in a mere


logical space, as was the case for Schelling, but in a real and historical
one. According to Treviranus, the «fundamental problem of
biology» is in fact the distribution of living bodies according to
different environmental pressures. Treviranus’ research on this
issue must start from the assumption «that all living forms are the
product of physical influences»26. Kielmeyer had stressed for the
first time the necessity of a general theory of animal organization
concerned with the laws that regulate the distribution of vital
functions in the animal kingdom27. This program had been devel-
oped by Schelling, who had envisaged Kielmeyer’s famous Rede as
the beginning of a new epoch of natural history28. Albeit with
some differences, they both aimed at the formulation of universal
laws capable of accounting for biological variety. These laws
involved the idea of nature as a self-regulating organism that
maintains its internal equilibrium through an equal distribution of
vital functions among the different species. This, however, does
not imply the idea of a dialectical relationship between organism
and environment. The emphasis on environment and external
conditions is what allows Treviranus to go beyond Schelling.
In his Ideen zu einer Philosophie der Geschichte der Menschenges-
chlechtes (1785) Herder had defended a theory of the transmuta-
tion of living species. This idea had been criticized by Kant as a
«daring adventure of reason»29. Treviranus embarked on this
adventure more than anyone else in the German-speaking world.
As with Lamarck, the engagement with the geographical distribu-
tion of organisms and the question concerning their ability to
adapt to the surrounding environment led Treviranus to the first
coherent statement of a theory of biological transformation. He

26 Ivi, p. 264.
27 A. GAMBAROTTO, Vital Forces and organization: Philosophy of nature and biology in
K.F. Kielmeyer, «Studies in History and Philosophy of Biological and Biomedical
Sciences», 48, 2014, pp. 12-20.
28 T. BACH, Biologie und Philosophie bei C.F. Kielmeyer und F.W.J. Schelling, From-

mann-holzboog, Stuttgart-Bad Cannstatt 2001.


29 I. KANT, Kritik der Urteilskraft, § 80, Akademie Ausgabe, 5, 419; cf. P. HUN-

EMAN, Naturalizing Purpose: From comparative anatomy to the “adventures of reason”,


«Studies in History of Biological and Biomedical Sciences», 37 (4), pp. 621-656.
150 Andrea Gambarotto Articles

begins by asking the following questions:

through which of these causes did living nature obtain the form
it has now? Did all the different genera of living bodies emerge
from formless matter, or only certain prototypes (Urformen),
while the rest arose from them through degeneration or for-
mation of bastards?30.

To answer these questions, the consideration of geographical


distribution is not sufficient anymore. Treviranus deems it neces-
sary to account for the modifications of living nature over time.
The purpose of the third volume of the Biologie is in fact to find
out «which transformation living nature went through before
obtaining its present formation (Bildung)»31.
It is possible to assume that

the organism of living nature, just as everything else in space


and time, is subject to infinite transformations. But then should
not the organization of living bodies change as well? Should not
entire kinds (Arten) perish and new one emerge?32.

Since a species (Gattung) cannot disappear from living nature


without effecting its organization, the downfall of a kind must
necessarily imply the emergence of another. Therefore, according
to Treviranus, for animals and plants that we register as “newly
found” in our indexes, the name of «newly produced» is perhaps
appropriate. These kinds (Arten) that were already present in the
first times of the history of the human kind, and that have re-
produced up to the present, are considerably different from their
previous form (Gestalt)33.
Treviranus maintains that nothing can be determined about
the history of living nature as long as we are uncertain about the
genesis and formation (Entstehung und Bildung) of the earth. For
this reason, part of the book is dedicated to a discussion of the

30 Ivi, p. 499.
31 G.R. TREVIRANUS, Biologie, oder Philosophie der lebenden Natur für Naturforscher
und Ärzte, III, Röwer, Göttingen 1805, p. 3.
32 Ivi, p. 21.
33 Ivi, p. 23.
Articles Teleology beyond Regrets 151

different layers of the earth. The oldest ones, which consist in


limestone, contain fossils of polyps and crustaceans only in small
number. Among them there are different kinds of slate contain-
ing not only remnants of vegetable-animal organisms (Tierpflan-
zen) and mollusks. The number of these organisms gradually
increases in rocks of more recent formation: in the following
layers, in fact, skeletons of fishes and invertebrate animals can be
found. These data suggest a modification of living nature in
which several of the previous kinds of marine animals progres-
sively disappeared while new ones emerged in their stead. In all
these layers there is no trace of land animals: bones of quadru-
peds can be found only in the most recent ones. These facts
imply that

the formation of living nature began from polyps and mollusks,


i.e. from the lowest levels of organization, progressing from
those to plants, and only afterward to land animals. A similar
process from the simple to the complex takes place today in the
generation of vegetable and animal substance from formless
matter in infusion34.

Living nature is to Treviranus «an eternally self-transforming


organism that progresses regularly towards a certain state of
development (zu einer gewisser Grade der Entwicklung)»35. In an infu-
sion, complex organisms develop from animal and vegetable
substance. If one considers that the whole living nature has also
gradually progressed from the simpler to the more complex, «so
it is clear that all life can reach the higher levels of organization
only from the lower». But how else could this not imply that
«simple organisms were progressively formed from generation to
generation?». All the remains of the prehistoric world «are the
original forms from which all the organisms of the higher classes
emerged through gradual development (Entwicklung)»36. From this,
he maintains, it seems to follow that the animals of the prehistor-
ic world were not destroyed by great catastrophes, but rather that

34 Ivi, p. 40.
35 Ivi, p. 173.
36 Ivi, 225.
152 Andrea Gambarotto Articles

«many of them survived and disappeared from the current nature


because the kinds to which they belonged completed the cycle of
their existence and transformed in other genera». Everything on
earth «is volatile and temporary, the kind as the individual, and
the genus as the kind. Even humans will maybe elapse and trans-
form». It is in fact possible to assume, not without reasons, «that
nature has not yet reached the highest level of organization in
humans, but rather that it will produce even more advanced and
elevated beings, more noble forms»37.
Treviranus’ theory of transformism can be understood as a
historical reinterpretation of Schelling’s notion of a logical succes-
sion in the graded series of organisms. Schelling’s philosophy of
nature partially constituted the metaphysical framework of the
emerging biology, but Treviranus’ emphasis on environmental
pressures allowed him to go beyond Schelling’s schema, interpret-
ing it in terms of a real transformation of living forms.

3. Conclusions

In conclusion, a careful analysis of the Biologie demonstrates


that we need to go beyond Lenoir’s regrets if we are to under-
stand the conceptual foundations of biology as it emerged at the
beginning of the nineteenth century. Scholars have already argued
that the alleged agreement between Kant and Blumenbach was
based on a substantial misunderstanding of the respective concep-
tion of teleology. In fact, Blumenbach ignores the Kantian distinc-
tion between constitutive and regulative principles and conceives
of the Bildungstrieb as a goal-directed drive proper to all organized
beings38. For this reason the Lenoir thesis can no longer serve as
point of departure for reconstructing the history of nineteenth-
century German biology. I especially agree with Zammito that the
«vital materialism» developed at Göttingen is not quite the Kantian
«transcendental philosophy of nature» that Lenoir wants it to have
been. On the contrary, we find the Göttingen School far closer to

37 Ivi, 226.
38 R.J. RICHARDS, Kant and Blumenbach on the Bildungstrieb, cit., pp. 30-32.
Articles Teleology beyond Regrets 153

the Naturphilosophen than Lenoir would like39.


The formalization of biology as a field implied in fact a con-
sideration of the teleological features displayed by living organ-
isms as their most proper characteristic, not as a mere imputation
of our faculty of judgment. This means that the formulation of a
general biology at the beginning of the nineteenth century was
the result of a conceptual shift from the Kantian conception of
teleology as a mere regulative principle, to a new understanding
of teleology as a constitutive property of living entities. This shift
is given a first philosophical formulation in Schelling’s Natur-
philosophie and was later endorsed by Hegel as well. Both consid-
ered teleology as an explanatory principle of a higher degree than
mechanism and chemical affinity, one that was necessary in order
to provide an adequate conceptual account of living organisms.
In Schelling’s organicist view, the teleological features displayed
by living organisms are understood as the link between the lower
plateaus of nature (the most extrinsic, manifested by the phe-
nomenon of mechanism) and higher ones (more intrinsic or
holistic, progressively manifested in chemical affinity and teleolo-
gy). Accordingly, ʻspiritʼ or the ʻmindʼ (Geist) are presented as the
final manifestation of this Stufenfolge.
I hope to have shown that Schelling’s organicist views are
frequently employed in key passages of the Biologie. In fact, the
Schellingian idea of nature as universal organism constitutes a
sort of metaphysical (or at least conceptual) foundation for
Treviranus’ idea of biology. At the same time, the emphasis on
the geographical distribution of organisms and the importance
granted to the environment projects the Biologie beyond this
framework. Schelling’s system of nature consists in the logical
deduction of an ideal series of organizations. Treviranus inter-
prets it as a real sequence, thereby providing the foundations for
biology as a historical science. This might lead us to the conclusion
that German Idealism played a relevant role in the formulation of
a general explanatory framework aimed at accounting for biologi-
cal organization. Only a historiographical bias, unsupported by
textual evidence, could lead us to think otherwise.

39 J. ZAMMITO, The Lenoir Thesis revisited, cit., p. 130.


THE CONCEPT OF ORGANISM
IN HEGEL’S PHILOSOPHY OF NATURE

by Luca Illetterati

Abstract. I focus my attention on the conceptualization Hegel offers of the


organism in his philosophy of nature. The aim of my paper is to show the
naturalistic roots of the notions of subject. Through this path I also intend to
shed light on the way the connections between these different notions – organ-
ism, subject, freedom - are capable of producing a certain re-definition and re-
determination of the immediate use of the terms with which these are usually
represented in ordinary language and the way they appear, prima facie, in
Hegel’s system. This process of conceptual re-definition and re-determination
of the terms that are here at stake could also be of some interest in relation
to the philosophical debate of these last decades on naturalism and anti-
naturalism. More specifically, it could shed light on the different ways of
inflecting the notion of naturalism in philosophical context.

At the beginning of the third part of the philosophy of na-


ture in the Encyclopaedia of the Philosophical Sciences (1830), Hegel
describes the organism as «an impregnated and negative unity, which
by relating itself to itself, has become essentially self-centred and subjec-
tive»1. To understand what these determinations constituting the
fundamental characteristic of the organism are, it is necessary to
look at that part of organic physics where Hegel discusses the
significance of the life of animals. Unlike rocks and plants,
where these characteristics are only formally or directly dis-
closed, but nor effectively and fully realized, it is with animals
that they are actually concretized.
The concept characterizing the animal sphere is, first of all,
that of subjectivity, a notion thematised for the first time in Hegel’s
systematic development of a naturalistic context. But in what

1 G.W.F. HEGEL, Enzyclopädie der philosophischen Wissenschaften im Grundrisse


(1830), GW, Bd. 20, hrsg. von W. Bonspien u. H.-C. Lucas, unter mitarbeit
von U. Rameil, 1992, § 337 (henceforth: Enz. C).
Verifiche XLII (1-4), 2014, 155-165.
156 Luca Illetterati Articles

sense is it possible to say that animal is subjectivity? What does


Hegel mean by stating that the animal’s way of being is a subjec-
tivity’s way of being? Animals are thus described in the 1817
Encyclopaedia:

Organic individuality exists as subjectivity insofar as its individ-


uality is not merely immediate actuality but also and to the same
extent suspended, exists as a concrete moment of generality,
and in its outward process the organism inwardly preserves the
unity of the self (die selbstische Sonne)2.

To understand these words, and especially what Hegel


means with the idea (which disappears in the English transla-
tion) that the animal, in its relation to the outside world, still has
a sort of selbstische Sonne – an image that summarizes on a repre-
sentative level the meaning it has in Hegel – it is necessary to
explain the way plants had been conceived: incomplete organ-
isms, characterized by a peculiar immediacy. Such immediacy
implies that on one hand, plants cannot be authentic unities
within difference. On the other hand, as plants have their de-
terminacy outside themselves, they revolve around something
else (the sun, or more generally, light) 3. What makes plants a
partial and immediate realisation of the concept of organism is
their specific characteristic that, in Hegel’s words, they have
another self outside themselves, an outside unity towards which they
tend and on which they depend. This self outside themselves is
primarily light, towards which plants turn, and that on them has
the strongest power. In fact, plants do not move of their own
accord, but are conditioned in their movements 4.
The main element of animal subjectivity is the negation of
such immediacy, appearing as a sort of liberation from the

2 Cf. G.W.F. HEGEL, Enzyclopädie der philosophischen Wissenschaften im Grundrisse


(1817), in GW, Bd. 6, § 273 (henceforth Enz. A).
3 Enz. A, § 269.
4 Hegel thus writes: «light is this physical element outside the plant towards

which it turns the same way man searches for other men» (G.W.F. Hegel, System
der Philosophie, in G.W.F. Hegel, Sämtliche Werke, Jubiläumsausgabe in 20 Bänden,
hrsg. von H. Glockner, Bd. 9, Frommann, Stuttgart 1929, § 344 Z, p. 500).
Articles The Concept of Organism in Hegel 157

dependency that characterises plants in their relation with


natural elements. The structure of an animal is such that the
target towards which it aims is not, as with plants, external. In-
stead, it identifies with itself5. Even when the organism’s activity,
starting from the need it is experiencing, moves away from its
singularity and towards what is other, it always realizes itself. This
means that the animal, in its inward activity, has a movement that,
in moving outward, always has in itself its objective and its centre.
This makes it a subject.
Since it has in itself its centre – the centre around which its ac-
tivity revolves, animal subjectivity is, according to Hegel, a concrete
unity. It is not simply a formal unity, as in plants, where the parts
are independent from the whole, and capable to keep on living
once severed from the whole giving birth to new consistent
wholes. That of animals is a concrete unity since it realizes through
difference and internal ramifications. It is a unity in which the parts
constitute the whole in a way that if they were separated from each
other, they would stop being what they are, losing their coherency.
The concrete unity of animal subjectivity is what makes animals
individuals in an actual and tangible sense, a way of being that
cannot be divided without being nullified in its ontological struc-
ture. Such structure is always one with itself, even in its internal
ramifications and always becoming other than itself6.
As subject, the animal has in itself the core of the principle
of its unity and thus differs greatly from both rocks and plants as
being the only one capable of self-movement. It is the only one
capable of not being under – even if only partially – the control

5 Karl Heinz Ilting and Franco Chiereghin have discussed this passsage from
plants to animals. See: K-H. ILTING, Hegels Philosophie des Organischen, in Hegel
und die Naturwissenschaften, ed.by in Michael J. Petry, Frommann-Holzboog
Stuttgart-Bad Cannstatt1987, pp. 349-368; F. CHIEREGHIN, Finalità e idea della
vita. La recezione hegeliana della teleologia di Kant, «Verifiche», 19 (1990), pp.
127-229. On the analogy of animal and sun see. G.W.F. Hegel, System der
Philosophie, § 350 Z, p. 576-577.
6 Animals are the concrete realisation of life in nature since «it is the one that

has all the parts in their freedom unites in it. It divides in it, gives them
universal life and sustains them in itself as their negative, their force» (G.W.F.
Hegel, System der Philosophie, § 342 Z, p. 491).
158 Luca Illetterati Articles

of exteriority (light, gravity, etc.) and to self-determine according


to its location, but also its own needs and reasons. It is not a case
then that in the very final section of organic physics the idea of
freedom appears for the first time. For Hegel, the concepts of
subject and freedom are deeply connected, to the point that the
two words are sometimes used to express one another. Animals’
subjectivity is expressed precisely in the capacity to free from the
necessary bond of the external forces that prevent the plant from
even the smallest form of self-determination (and thus freedom).
Hegel connects and explains the animal’s possibility and ca-
pability to change its dwelling place as the peculiar relation that it
has with time. If the plant has to rely on light, especially when it
comes to its movements, it is also dependent on nature’s cyclical
passing of time for its growth, nutrition and reproduction. Ani-
mals instead, require what Hegel calls «free time»7. This expression
means that animals are, to a certain extent, independent from the
external and natural time to which plants are subjected, which
makes them autonomous and capable of self-determination. This free
time manifests itself through self-movement, which cannot be
merely understood as moving from one place to another. It is
«ideal» self-movement, a condition that is origin to all those
characteristics that define the animal way of being and that
constitute the particular determinations that will eventually find
new development at the level of spirit. These are the vocal
faculties, animal heat, the interrupted intussusception, and, above all
else feeling (Gefühl).
The vocal faculty is, for Hegel, the organism’s expression of
«free vibration within itself»8 and in this sense expression of its
subjectivity. Surely, the Stimme that characterizes animal subjectivi-
ty is not yet concretized in the symbolic production that will be
recognized, at spirit’s level, of actual language. However, the
Stimme, as manifestation of the animal’s subjectivity in its expres-
sion, exteriorization of its interiority, pain, satisfaction or feelings,
can be read as a sort of natural precondition to that symbolic
ability that will develop only at the level of spirit. Vocality is not

7 Enz. C, § 351.
8 Ibidem.
Articles The Concept of Organism in Hegel 159

simply the consequence of some internal mechanism of the


organism. Since it is exteriority of self-movement, it is self-
production, a phenomenon through which animals express their
self to give a form to their subjectivity and to their Gefühl9. Only
because the animal feels, it can express through its voice what
could be called, without necessarily implying self-consciousness,
its Self. Gefühl constitutes the determination through which the
animal feels itself, its own self, submerged in pain, pleasure,
satisfaction or suffering, in all situations which the Stimme can
exteriorize and objectify.
The animal’s subjective structure is further expressed by Hegel
in the context of the relation with the outside world, which is an
assimilative process. This relation begins through subjective feeling,
connected to the self. And the first feeling is loss. Thus, animal
subjectivity develops as the «push to supress»10 such sense of loss.
The assimilative process starts from a specific need determined
by a structural deficiency, and by possibility, a characteristic found
only in living beings and that determines their intimately subjec-
tive structure to feel such need and deficiency11. Deficiency, need,
intended as the perception of such deficiency, and the push to
satisfy that need are fundamental elements in Hegel’s conception
of living beings and natural subjectivity. The living being first of
all has a need, which is an integral aspect of its essence. This
means that if a living being did not have needs or deficiencies, it
would not be a living being anymore. Any living organism, no
matter its size or complexity, needs to demolish and rebuild its
constitutive materials through its metabolic activities: assimilation,
transformation and elimination. Being in need is the way a living
being exists, and through the processes of transformation and

9 Ibidem.
10 Enz. C, § 359.
11 «Only living being feel loss» (Enz. ‘30, § 359 An.). In The Science of Logic Hegel

writes: [G.W.F. Hegel,Wissenschaft der Logik. Zweiter Band. Die subjektive Logik
(1816), Gesammelte Werke, Bd. 12, hrsg. von F. Hogemann u. W. Jaeschke,
Meiner, Hamburg 1981, pp. 187-188 (trad. it. p. 874)]. See also G.W.F. Hegel,
System der Philosophie, § 358 Z, p. 632. It is important to underline that pain is
not the same thing as loss – otherwise it would not be a living being privilege.
It is the capacity to feel it.
160 Luca Illetterati Articles

modification takes in what is other from it, using for its own
construction of what is external.
A living being is in constant transformation, in a process in
which the organism acts on itself and on the outside world in
order to continue being in transformation, to keep on being itself.
This being in constant need in order to exist (die Tätigkeit des
Mangels) is what differentiates living from inorganic matter, which
is always the same and does not have any constitutive
lack12.Saying that the living organism is marked by its need does
not mean saying that it needs something else to be considered a
whole. Therefore an organism needs something to be itself the
same way a car needs gas. A living being is a process and it never
stays the same. If two stages of this process were absolutely
identical we could say that the being has ceased on living. Howev-
er, it can still be defined as a system that is always a unitary
whole13. Thus, deficiency is not simply a weakness that can be
overcome, or realizing that there is a missing piece that prevents
the system from working. Deficiency is integral to life. If it is true
that we consider complete beings that are complete vìs-a-vìs their
constitution, and that life is acting on a deficiency, what life needs
is need itself. Without it, life would not be life14. Deficiency and
need cannot be understood as defects or interruptions that can
be solved to gain constant fulfilment. Life’s peculiarities and
potentialities are not different from the negativity of need. They
are entangled in this way of being.

12 The expression activity of lacking (Thätigkeit des Mangels) is used by Hegel to


determine the structure of impulse (Trieb) belonging to living being. See G.W.F.
Hegel, Zum Mechanismus, Chemismus, Organismus und Erkennen, in Gesammelte
Werke, Bd. 12, hrsg. v. F. Hogemann u. W. Jaeschke, Meiner, Hamburg 1985,
pp. 259-298, in part. p. 280. In relation to this text and its value for theory in
general and Hegel specifically, see the Italian edition G.W.F. Hegel, Sul
meccanismo, il chimismo, l’organismo e il conoscere, trad. it. L. Illetterati introduction
and comments, Trento 1996, p. 54.
13 See H.R. MATURANA - F.J. VARELA, Autopoiesis and Cognition. The realization of

the Living, Kluwer, Dordrecht 1980.


14 To clarify the many meanings of neceessary, Aristotle says «necessary means

what it is impossible to live without». Breathing, eating. (ARISTOT., Metaph., V,


1015 a 20 sgg.). But since need, food and air are a form of deficiency, it can be
said that lack itself is necessary.
Articles The Concept of Organism in Hegel 161

Animals, then do not simply lack something, but they also


live and experience this deficiency within themselves. It is because
of this feeling of lacking something, and the consequent inherent
contradiction and pain, that the living being is the real subject.

The subject is a term such as this, which is able to contain and


support its own contradiction; it is this which constitutes its infini-
tude (Enz. ‘30, § 359 An.).

The infinity connected to the subject in the passage above has to


be understood as the possibility it has to let go of the concrete
shapes of need and deficiency. The subject’s infinity is its capacity
to perceive its contingency, to express its negativity, to live its
limit and to push it. It is thus revealed how the subject can trans-
cend itself the very moment it is determined as limited. Thus the
subject’s essential finiteness, its limitation and structural insuffi-
ciency emerge as biological conditions. The tension the organism
experiences to overcome its condition, to pass the limit, to satisfy
its restlessness pushes it to engage with the outside world, and
makes it what it really is.
In this sort of double process, where animal subjectivity per-
ceives itself and finite, and transcends its limits, only to discover
itself, once again, as finite, is particularly evident in Hegel’s analy-
sis of sexual relations and reproduction. In reproduction and
sexual relations, individuality opens to the outside world in the
hope of finding in another individual the completeness it lacks, to
integrate, through this union, its ontological weakness, and “to
bring the genus into existence by linking itself into it”15.

15 Enz. ‘30, § 369. On similarities and differences between gender, Hegel insists
in the 1805/06 Jena Naturphilosophie where he analyses sexual organs and
quotes specific researches such as J. F. Ackermann’s and G. H. Schubert’s (see
JS III, pp. 173-174). It is possible to see a correspondence between men’s
testicles and women’s ovaries, for instance, but beside all the possible
analogies, there is an essential difference, whereby the female is characterized
by being indifferent and the male instead by opposition and by the division,
from which follows that the male is the active element, the bearer of the
principle of subjectivity, while the female is receptive, the matter must take the
form (see JS III, p. 173-174). The reference to the ancient Aristotelian theory,
according to which the male provides the form and principle of change (archén
162 Luca Illetterati Articles

The other individual shares the same sense of deficiency,


fragility and insufficiency («that feeling of insecurity» 16, Hegel
says) that pushed it to look outside. However, this attempt is
inevitably a desperate one. Unlike what is described in Aristopha-
nes’ tale in Plato’s Symposium, sexual relations are not the integra-
tion and mutual fulfilment of two finite and isolated entities.
Rather, they are the reason for the birth of a new individual, a
new singularity that has the same feeling of deficiency and onto-
logical inadequacy as the other two. The attempt to overcome
such inadequacy is both reason and origin of its existence. The
individual’s struggle is solved in nature with that bad infinity to
which the individual is destined to succumb:

This process of propagation issues forth into the progress of


the spurious infinite. The genus preserves itself only through
the perishing of the individuals, which fulfil their determination
in the process of generation, and in so far as they have no high-
er determination than this, pass on to death17.

The genus exists only through the death of the individual, and
thus is a higher form of life than the single entity, which is always
divided in its universality. It is a natural form of life that however,
sometimes, also transcends nature:

In this new life, in which singularity is removed, subjectivity is


maintained, and the genus has become, for itself, reality, becom-
ing something higher than nature18.

tés kinéseos), while the female the body and matter, is obvious here (ARISTOT.,
De generat. 1, 729a). The reference to Ackermann, who taught anatomy at Jena
in 1804, is not devoid of interest because his works were probably a significant
influence in the scientific training of Hegel. Ackermann had already published
in 1806 a work in which he undertook to show the unsustainability, from a
scientific point of view, of the phrenology of Gall against which Hegel wrote
against at the same time in the Phenomenology of Spirit. Ackermann’s work,
published in Heidelberg in 1806, is entitled Die Gall’sche-Hirn, Skull-, Organ-
und Lehre vom Gesichtspunkt der Erfahrung Enz.
16 Enz. ‘30, § 369, An.
17 Enz. C, § 370.
18 Enz. A, § 291.
Articles The Concept of Organism in Hegel 163

Spirit is what is higher than nature. Here, a reconfiguration be-


tween individual and universality occurs. This reconfiguration is
gradual and it is never fully free from objectification, apart from
the moments of complete awareness. Nature and the outside world
that it embodies, and the tear of deficiency that it manifests in its
most complex form, do not disappear in this reconfiguration. It
gains a new and different meaning that reorganizes and gives new
structure to that very same exteriority, deficiency, and need. In
animal subjectivity nature – which is primarily exteriority – is
fulfilled. Here nature reveals its conceptual structure that, in all its
other manifestations, was always only internal and separated from
any objectivity. If fulfilment is acknowledgement and revelation
of what it really is, nature, through animal subjectivity, reveals a
peculiar tendency to go beyond nature itself and that strict neces-
sity that, according to Hegel, is a necessary characteristic of
nature and being other. What is interesting is that this movement
to overcome this strict law of nature does not act from the out-
side. It is in nature itself, thus allowing and making necessary a
redefinition of the concept of nature itself.
The broader conception of nature that makes Hegel think
about the relation between nature and spirit as different, but
never opposed world, does not seem to be unrelated to nature’s
essence. The structure of subjectivity and the consequential
freedom are not the outcome of some kind of infection of spirit
on nature, or of an external influence that initiates something
that would otherwise remain unscathed from this type of dynam-
ics. Life is a manifestation of nature. The structure of subjectivity
and the freedom that exists in it are nature’s highest achievement
in terms of organization and structure. According to Hegel, the
limitations of physical reductionism (and of strict naturalism)
do not appear out of anti-naturalistic assumptions, but from the
radical consideration of nature’s essence. From a certain per-
spective, Hegel’s position seems, on one hand, to go towards a
naturalization of the subject, showing how the subject’s way of
being (the subject is intended here as a structure revolving
around itself, autonomous and self-determined) develops pri-
marily in nature and, specifically, in animals. On the other hand,
however, it also involves a redetermination of the idea of nature
164 Luca Illetterati Articles

with a process that can be seen as sort of denaturalization of


nature, and that Hegel would describe as unilateral, intellectualis-
tic and reductionist.
Finding the genesis of subjectivity in nature prevents from
thinking about it as a disjunctive element, as something that
would appear only after nature and within the social practices and
dynamics connected to it19, or as the bursting in of a supernatural
principle on a natural layer. However, understanding nature as the
place where the subject literally takes shape prevents seeing it as
simple exteriority with no freedom, the way in which, at least
prima facie, it is constituted within a systematic structure. Thinking
about the subject and about freedom in a radically naturalistic way
prevents seeing nature and spirit as juxtaposed, as if opposing a
determined-by-necessity nature with an independent supernatural
reality. The contraposition between nature and spirit starts, in-
stead, from abstract conceptions of both notions. Through this
process of conceptual redefinition aimed at overcoming intellec-
tualistic abstractions, Hegel attempts to show spirit’s development
in nature and nature’s redefinition in spirit. In this perspective,
second nature is not only erasing first nature – what Hegel would
have called natural nature – but it is a new redefinition of the
complex human structure, of the subjective structure of man as
an organism.
Second nature, the grounds on which the human way of being
and spiritual world develop, is rooted in human being’s free sub-
jectivity, in his being a development of those characteristics that
essentially define animals as such. Hegel aims at solving any form
of dualism characterising some of the relations with the outside
world and that are the origin of a certain spiritual and physical

19 The argument here highlights the limits of the interpretations of Hegel’s


philosophy, which emphasized the social dimension as the original place where
the structures of subjectivity and freedom are revealed. It is in many ways
around this problem that the controversy between J. McDowell and Robert
Pippin develops. Nature Leaving behind that Pippin wrote against McDowell
implies a conception of subjectivity and freedom in Hegel that is intended to
show the elements that break nature and that are irreducible to any form of
rationalism. Equally apparent in Pippin in his polemic against DeVries’
emergentist Hegel.
Articles The Concept of Organism in Hegel 165

reductionism. For Hegel, spirit is not simply something different


from nature. This dichotomy, to use Wittgenstein’s terms, is a
classic conceptual pathology. Spirit cannot appear unless the
natural bonds where it originates and develops are recognized.
And if spirit is not different from nature, since it arises from
human beings’ nature, it is clear that such condition necessitates a
further development of the concept of nature.
The opposition to a physicalist reduction of nature does not
produce a spiritualistic ontology, nor a reduction of reality to the
mind, as in a classical but radically idealistic reading. Materialism
and spiritualism have sense only within the abstract and opposing
logic that maintains them. They are unilateral determinations that,
in the overlaying dualistic vision, are each other’s reversal. The
appearance of subjectivity within nature, and the decline of animal
subjectivity through relations that require freedom is proof of the
need to let go (also in a therapeutic sense) of all the dualisms and
abstractions that are at the origin of many forms of reductionism.
This need is the fulfilment of Hegel’s system in its divisions as
logic, philosophy of nature and philosophy of spirit, and its
development as a whole in which every part makes sense only in
relation with the others and with the whole.
BOOK REVIEWS

S. NORMANDIN, CHARLES T. WOLFE (eds.), Vitalism and the Scien-


tific Image in Post-Enlightenment Life Science, 1800-2010, Springer,
Dordrecht 2013, 377 pages, ISBN 978-94-007-2445-7.

This volume consists of a collection of essays on the history


and philosophy of vitalism. The volume is edited by two re-
nowned experts, Sebastian Normandin and Charles Wolfe. The
latter especially can be considered as a leading scholar on the
subject, having written several important papers on related topics
and having been editor of a volume entirely dedicated to medical
vitalism in the Enlightenment. Compared to that previous collec-
tion of essays, this volume is much more ambitious. In fact, it
embraces a broader perspective, undertaking «a history of vitalism
in the era of synthetic life». As Wolfe has stressed in previous
studies, vitalism is less monolithic and less doctrinaire than the
general idea conveyed with this label. It has been a category
applied to various theories that have little in common with each
other with entirely different empirical bases and metaphysical
commitments. Here vitalism is portrayed as a «moving target»,
«an explanatory and/or metaphysical construct which appears,
depending on the context, as a form of overt supernaturalism or
as useful heuristic for biomedical research and theorizing». Un-
derstood in this manner, vitalism is far from dead: contemporary
debates on holism, emergence and complexity testify to a growing
interest in ʻvitalisticʼ stances. Using Canguilhem’s words, vitalism
rather has its own vitality.
The volume is divided into three different sections: I. takes
into account vitalist themes in nineteenth-century science, II.
deals with twentieth century debates on vitalism in science and
philosophy, III. considers the role of vitalistic stances in contem-
porary biological developments.
I. As is well known, the pivotal role in nineteenth-century vi-
talism is played by physiology. This section is consistently dedi-
Verifiche XLII (1-4), 2013, pp. 167-188.
168 Book Reviews

cated to show the developments of the Hallerian framework in


France, Germany and England. The chapter by Guido Giglioni
shows how Lamarck appropriated the notion of irritability in the
revised version popularized by Haller, which had an important
tradition in France, including Bordeu, Maupertuis, Diderot and
Barthez. By espousing the notion of irritability Lamarck joined
this long-established tradition of medical thought while at the
same time providing an original reinterpretation of the phenome-
non: he rejected the hylozoistic implications while retaining its
explanatory potential with regard to organic change. In particular,
he used the process of irritability to account for various mecha-
nisms of life-adaptation and self-preservation.
Juan Rigoli explores the circulation of vitalist stances in mid-
nineteenth-century France, when physiology became a conquer-
ing science which laid claims to exclusively describing the entirety
of the human experience. He focuses especially on Claude Ber-
nard and gives great attention to relevant literary “vulgarizations”
by Balzac, Zola and Flaubert, which testify the importance of
vitalist concepts even beyond the practice of physicians. Comply-
ing with the old ʻLenoir thesisʼ, Joan Steigerwald focuses on the
key-concept of German physiology, that of Lebenskraft, to argue
that vitalism in Germany around 1800 was not a product of
speculative philosophies of nature. She takes into account the
vivid debate involving Blumenbach, Girtanner, Kielmeyer, Hum-
boldt, and Reil, which led to the formulation of the idea of a
general biology by Roose, Burdach and Treviranus. She considers
Schelling and Oken as well, to show the multiple solutions pro-
vided to the problem of organic vitality at that time. Sean Dyde
focuses on the theories of life and mind in nineteenth century
Britain. He shows how Combe’s Phrenology, Hall’s theory of
reflex action and Laycock’s theory of cerebral reflex function
aimed for a physiological account of nerves while fumbling to-
wards a vitalistic theory of mind.
II. The major point of interest of the second section lies in
the focus on the notion of «emergence». Christophe Malaterre
traces the long history of the concept, through golden ages,
setbacks and comebacks. Since its early philosophical formulation
in the nineteenth century by John Stuart Mill and Henry Lewes,
Book Reviews 169

the notion of «emergence» has been applied to vital phenomena,


but also to chemical compounds and mental states. In each case
the whole is said to be more than the sum of its parts, as higher
levels of organization appear to exhibit properties that are
claimed to be non-deducible, non-predictable or unexplainable on
the basis of the properties of its lower level of components. In
the early twentieth century the notion reached its golden age in
the philosophical writings of Cowny Loyd Morgan and Charlie
Dunbar Broad, being construed as an alternative to vitalism and
its dualist ontology, but also as an alternative to mechanism and
its constraining determinism. The concept encountered several
scientific and philosophical setbacks in the mid-twentieth century
with the advent of quantum mechanics and the simultaneous rise
of molecular biology and prebiotic chemistry, as well as with the
criticism in philosophy of science in the wake of logical positiv-
ism. It somehow re-emerged in the late twentieth century becom-
ing a central topic in the philosophy of mind and receiving unex-
pected support by the science of complex systems. Even
nowadays its relevance remains a debated topic.
The essay by Brian Garrett articulates the fundamental tenets
and aspirations of vitalists and emergent evolutionists between
1900 and 1930. He stresses the broad similarity between those
doctrines and those discussed in recent debates. Although the
stage has changed and the ʻpropsʼ are more sophisticated, he
argues, the core problems and arguments remain pretty much the
same. He thus suspects that the debates of the last thirty years
could have been moved faster if scholars had not been affected
by a serious case of «intellectual amnesia». The essays by Sebas-
tian Normandin, Chiara Ferrario & Luigi Corsi, and Giuseppe
Bianco present specific ʻcase studiesʼ of peculiar figures of twen-
tieth-century vitalism: Wilhelm Reich, Kurt Goldstein, and
George Canguilhem.
III. The third section is entirely dedicated to the role played
by vitalist stances in contemporary biological debates. The essay
by Scott Turner argues against the reductionist tendency to con-
sider properties of living systems, including purposefulness,
design and intentionality as illusions. Those phenomena, he
maintains, are far from illusory: they are, in fact, quite real and
170 Book Reviews

should be accounted for in a coherent theory of biology. This is


possible, he believes, through a consideration of homeostasis.
The essay, however, focuses especially on Claude Bernard’s
notion of milieu intérieur and one feels the lack of a proper argu-
ment for these broad philosophical claims. Carlos Sonnenschein
et al. discuss the beginning of tissue culture starting in the period
1907-1912. They thereby address the unchallenged concepts and
the unasked questions behind the experiments of tissue culture
pioneers, who assumed quiescence instead of proliferation as the
default state of the cells in metazoans. This assumption required
the hypothesis of a so-called «growth factors» that would stimu-
late passively quiescent cells to undergo proliferation, and implied
the existence of an external, undefined entity instructing the cells
to enter their cycle of reproduction. This mistaken conclusion,
they argue, exemplifies the consequences of the naïve physicalism
that hinders the understanding of biological organization.
John Dupré and Maureen O’ Malley address issues related to
the definition of life. Scholars have widely debated two opposed
views: the first pointing at the capacity to form lineages by repli-
cations, the second to the capacity of existing as metabolically
self-sustained wholes as the core-character of life. According to
the authors this tension can be best resolved by seeing life as
something that arises at the intersection of these two features:
matter is living when lineage-forming entities are (directly or
indirectly) implied in metabolism. To testify the open and hospi-
table attitude of this collection of essays is the presence of one of
the most renowned contemporary ʻmechanistsʼ, William Bechtel.
He argues that vitalists correctly objected that mechanists’ expla-
nations in biology lacked the resources to explain important
features of biological phenomena. In his view, the key to address-
ing these objections is to incorporate those features within a
mechanistic explanatory framework. In particular, it is necessary
to understand how non-sequential organization enables mecha-
nisms the sort of complex behavior exhibited by living organisms.
This requires developing mathematical equations to represent
these operations and computational simulations to determine
how various components of the mechanism change depending on
their own state and those of other components.
Book Reviews 171

Vitalism and the Scientific Image is an ambitious book which


provides interesting insights for the contemporary philosophy of
biology. It especially makes an important point in showing that
conceptual history is a useful tool to frame our contemporary
problems correctly and to avoid new cases of intellectual amnesia.

(Andrea Gambarotto)

FRANCESCA MICHELINI, JONATHAN DAVIES (eds.), Frontiere della


Biologia: prospettive filosofiche sulle scienze della vita, Mimesis, Milano
2014, 328 pages, ISBN 978-88-575-1784-1.

The volume «Frontiere della Biologia» was published in 2013


by Mimesis under the editorial direction of Francesca Michelini
and Jonathan Davies. All contributions are in Italian and are
elaborated from papers presented in the international seminar
«Frontiere della Biologia», held at Fondazione Kessler in Trento.
The book declares its intentions upfront, starting with a most
peculiar title: «Frontiere» (i.e. «boundaries»), which has in fact an
often-unnoticed ambivalent meaning. On the one hand, it indicates
the limits within which the object in question is exhausted, but on
the other it also embodies the space where two separated areas
«touch» and mix one into the other. A boundary could hence mean
a passage to beyond-regions open to exploration, signifying a
difference and a connection at the same time.
Concentrating on this duplicity endorses a perspective of «uni-
ty in diversity», where a unitary insight leads to methodological
definitions and programmatic agendas.
The volume intends to inquire the «boundaries» of biology in
both senses of the term: it aims at a methodological redefinition
of biology as a science while it points at new directions for bio-
logical research and interdisciplinary confrontation.
The book hence draws the future of biology in two different
ways: while it indicates its furthest frontiers and objectives, it
shows how biology itself could be the most efficient mean to see
interrelations between all other natural sciences, turning tables to
172 Book Reviews

the reigning paradigm of hyper-specialism in favor of a renewed


conception of natural inquiry as a whole.
Biology is then meant to be the perspective from which
boundaries are seen as points of contact, rather than lines of
sharp demarcation. This is, according to all contributors, the main
by-product of the «revolution», which occurred in the world of
natural inquiry since the discovery of DNA, which at its turn
signified the rise of the «century of biology» and the end of the
times in which physics was considered the core science instead.
Biology is now experiencing an enhancement of attention
and importance of its research on the one hand and a broadening
of its objectives on the other; thus, not only a redefinition of its
methodological basics and an establishment of its programmatic
agenda is urgent, but also a meta inquiry aimed at reconstructing
the interdisciplinary changes that the «biological revolution» could
carry along. Here is where philosophy comes into play, question-
ing the consistency of epistemological frameworks and ontolog-
ical implications of naturalistic description, as well as its own
relation to biology, possibly falling under the «boundary para-
dox» while embodying it so well in its constitutive questioning
of foundations (i.e. in considering the concept of observation,
bioethics proves itself impossible to be considered as an area of
philosophy as sharply separated by all others, such as epistemolo-
gy and bio-philosophy).
The volume is a collection of fourteen essays from various au-
thors, all of them unpublished with the exception of two (Dupré-
O’Malley; Griffiths-Stotz), accompanied by an introduction by the
editors. The essays are grouped in three sections of interest, corre-
sponding to the triple aim of philosophical reconsideration of biolo-
gy and natural sciences in general: philosophy of nature, biology and
philosophy and epistemology of science. Volume partitions are not
meant as sharply separated from each other, and they acquire their
ultimate relevance only in the framework of an analysis of methodo-
logical support.
Every essay focuses on one keyword, expressing a funda-
mental concept in biological inquiry, and outlines a state of the
art summation of research advancements on the matter, its po-
tential directions of development and how it is through these very
Book Reviews 173

two considerations that the urgency of a new perspective for


interdisciplinary confrontation is shown.
The first part, «Philosophy of Nature», re-examines classic
matters of the philosophy of life, such as what difference, if any,
there is between man and animal, what defines human nature in
its specific character and what defines life in itself, with the relat-
ed question of the relation between life and death and whether
aging is an intrinsic natural process or an evolutionary selected
variant.
All essays of the section («Life», Dupré-O’Malley; «Human
Nature», Rasini; «Animal», Costa; «Aging», Giaimo-Boniolo;
«Environment», Gagliasso) cope in different ways with the pecu-
liar notion of «environment», which is thematically analyzed in its
new perspectives and meanings by the last essay in the group.
That is also the ground notion, which permits the recovery of
long-forgotten questions from classic natural science and philos-
ophy of nature.
While in fact recent research in biology tackles the common
concept of life as referred to cellular-based entities and shows
how biological identity is instead founded in original collabora-
tions between organisms, proving the relevance of internal envi-
ronment to the grounding of apparently independent individuali-
ty, it is the external ambient which is important for the definition
of human nature, the one stressed by a close analysis of the no-
tion itself. Likewise, a flip between conceptions of identity, indi-
viduality and relation or interaction is produced in the analysis of
the notions of animalitas and aging: while a peculiar overlapping
of the roles of the observer and the observatum lies in the meta-
physical mystery of our own animal component as humans, aging
is becoming more and more an enigma due to recent research on
the matter, challenging its common concept as an inescapable
fact and introducing the possibility of its being a product of
natural selection and as such, again, not an intrinsic character of
individuals, but an environment-bound feature.
All those revolutions and changes in scientific common
sense make a specific question on the notion of «environment». Is
«environment» a non-problematic notion in biology? Does it have
one meaning, or is it in itself polysemous?
174 Book Reviews

Of course, the answer to both questions is no, but this, in-


stead of constituting an obstacle, is exactly what permits this
notion to be the starting point for the construction of new path-
breaking approaches to what is apparently already known. Until
recently, the general tendency towards the notion of environment
was in fact the one of proceeding to a rigid semantic separation
according to the context of usage; there are distinct meanings of
«environment» depending on the discipline of reference, for
instance there is a notion of environment in ecology which is
considered to be completely detached and unrelated to the one
given in biology and micro-biology. While environment could then
define limits between different areas of research, it could constitute
a boundary between them also in the sense of a space of intercon-
nection and therefore as a ground of interdisciplinary confronta-
tion. Such a consideration opens to a meta-dimension of renewed
inquiry in epistemological, historical and philosophical matters on
the one hand, and endorses new connections between biology,
micro-biology and ecology on the other. While the new frontiers of
discovery in biology urge in the direction of eco-evo-devo theories,
stressing the intrinsic connection between evolution, natural selec-
tion and environment-based changes in organisms, both external-
ly and internally, would endorse an all-inclusive notion of envi-
ronment, which would follow what was considered one of the
two main merits of the Darwinian revolution, that is, a concep-
tion of nature as a whole in the process, a «Bioma».
It was from this seed that ecology evolved through Lamarck’s
heritage and nineteenth century botanic insights, becoming an
independent discipline, though forgetting its binds to biology,
which was even meant by some to become only one of its parts in
the consideration of the totality of its conditions of existence.
It is only in the renewed attention to means of natural selec-
tion other than DNA in biology that a rethinking of the ecology-
biology relation is found, and exactly along the lines of the similari-
ty of their changing and opening concepts of environment.
In this sense, not only the notion of environment can as-
sume many directions for further research in (eco)biology, but it
also endorses a change of perspective in epistemological attitude:
it enhances the perception of an organic whole not only in se-
Book Reviews 175

mantic definition and methodology, but most importantly in


nature itself, which is discovered as an active process of constant
selection and change, and relationally-based at its core.
The section then builds on the environment bind in two main
meanings: while it declines different shades of the peculiar flip
between an individuality-based consideration and a relationally-based
one in natural inquiry, it indicates ways to it in opening to influence
between the different sciences and their fields of research.
The second section is «Biology and Philosophy», and, as the
name suggests, focuses on the interaction between philosophical
inquiry and biological research and its potential benefits. The
section builds on a specific conception of philosophy, namely
that of a science of semantic clarification, which makes all the
contributions of this part more specific on the relationship be-
tween so-intended theoretical speculation and natural science
compared to the preceding group of essays.
All contributions reconstruct a genealogy of their keyword of
interest, showing that it is by historical semantic recollection and
definition that different methodological hints and epistemological
questions emerge to fit together with current research. In this
sense, philosophy as a work of clarification on texts and words
plays a preliminary role to all research, ensuring not only its
historical accuracy and a sound confrontation with tradition, but
also conceptual richness and correctness in the acknowledgment
of metaphysical frameworks. While a semantic study of the term
«Darwinism» (Depew, Weber) enables one to see the necessity of
an extension of the term’s meaning and henceforth of its theoret-
ic ground in order to enclose new dynamic models of evolution
(some of which were for instance named and briefly presented in
the previous essay on environment); the inner controversial
character of the term «Species» (Continenza) suggests a re-
examination of the ter’s implications with matters of the history
of science in order to lift the clash between topological and popu-
lation-based approaches.
Philosophy helps make distinctions preliminarily to applica-
tive research, as in the case of the term «Gene» (Griffiths, Stotz),
where three different meanings need to be kept sharply separated
(instrumental, post-genomic, nominal); furthermore it can show
176 Book Reviews

how some concepts not commonly related with biology fit with
its newest results and discoveries, giving them proper description
and conceptual classification, which happened with the term
«Network» (Civello).
It is down both lines, prior to applicative research and a poste-
riori that the essay on the concept of «Organism» (Toepfer) is built
upon. While it draws a genealogy of the term in its revolutionary
effect towards mechanistic, or Cartesian, conception of nature, the
essay shows how recent developments of research require a second
analysis of the term, in order to differentiate between a functional,
broader, meaning and an operational one, so to avoid metaphysical
misunderstanding and relativistic drifts.
Originally, the conception of «Organism» was connected with
the one of reciprocity, both between parts composing the whole
(i.e. organs) and between different functions and enacted process-
es. In this sense, organism was an immanent notion aimed at
replacing the transcendent, non-scientific one of «soul», starting to
take over with the Enlightenment. «Organism» was hence intended
as intrinsically linked to the notions of «life» and «living being». As
a mean of immanent unity and activity, the notion of organism was
run by a causal cycle securing autonomy: that is, in an organism
every function is also a goal as a principle of its autonomy. Teleol-
ogy became then the dominating perspective in the study of organ-
isms and in the definition of the term, as best brought up by Kant
basing on eighteenth century botanic legacy. Teleological consider-
ation took over mechanistic conceptions since it was able to ac-
count for a scientific grant on intrinsic unity amongst the parts.
Such a unitary flare was abandoned in biology when the discovery
of genes and DNA gave considerable push towards an ʻatomisticʼ
conception of living beings, where life was intended to be truly
embodied in cells and not in organisms and, most especially, as a
process so much extended over time, as over the time window of
advancement processing in evolution, that organisms could be
nothing but transient, irrelevant and partial embodiments.
If recent research in biology and ecology conducted a fairly
diffused renaissance of the concept of organism, there are several
metaphysical side effects and traps that come with this return in
fashion. The traditional conception of the organism as a func-
Book Reviews 177

tional unity produces in fact several misunderstandings, when this


is applied to recent research outcomes on internal and external
environment based relations and to the implications they had on
biological and morphological identity. It is in fact unclear if the
functional definition of organism applies to all functional aggre-
gates, also to non-living ones, and, most especially, how does
such a functional definition apply to organism-environment
interactions. A functional consideration of the organism might in
fact lead to an overestimation of the interconnection between
organism and environment, and lead to a naïve pantheistic con-
ception, in which the boundaries between the commonly meant
single living being and its habitat are completely erased due to the
fact that the environment too is participating in the functional
process defining the organism as one entity. While this might be
open to a biological ground for the speculation of nineteenth
century philosophical anthropology (Plessner and Scheler), it
could be risky to follow the slippery slope of holistic simplifica-
tion. It is then contemporary philosophy of nature and bio-
philosophy's duty to preserve the polysemous character of the
notion of organism, insisting on its controversial edges to keep it
linked to the enigmatic character of «emergence», which made it
the core notion of life sciences. The essay then calls the attention
to many apparent counterintuitive features of the term: for in-
stance, while it is commonly connected to the notion of a living
being as its innermost character, «organism» cannot coincide with
the notion of «body», since the latter lacks in functional determi-
nation and covers exclusively (in most cases) its morphological
determination; on the other hand, another controversial aspect of
the functional definition of the organism is posed by reproduction
and death: how relevant are these to be considered if the meaning
of the organism is to be built on its functions? What would this
imply on the morphological side then?
The third and final group of essays goes under the title of the
epistemology of science. The section tracks all the main episte-
mological side effects carried along by the recent biological revo-
lution, that is the substitution of biology to physics as the leading
science in naturalistic explanation. It is on this very train of
thought that possibilities of shedding light on new interactions
178 Book Reviews

between research fields are analyzed and linked to epistemological


concerns that might arise. Hardly surprising, the section proves
how a by-product of this very epistemic revolution is to have
brought up several old concerns in the philosophy of nature,
forgotten at once with the physicalist paradigm take-over in
nineteenth century. In a mix between innovation and tradition,
the section tackles common sense on epistemic frameworks,
showing how a twist in the past might suggest an ontological
enhancement of those patterns, to better face the challenges of
the future of research.
While this fil rouge is declined on the three classic models of
explanation in «Mechanism» (Powell), «Emergence» (Davies) and
«Teleology» (Michelini), the opening essay of the section, «Obser-
vation» (Köchy) puts the epistemic point under a rather more
general and at the same time more radical perspective.
The act of observing is in fact the basic action towards which
epistemology is oriented, questioning the notion then implies a
reconsideration of epistemology in itself and from within. The essay
moves from Merleau-Ponty’s suggestion of a relational take on the
act of observation, endorsing a dimension where a determined
epistemic attitude corresponds to a different onto-logical grasp. In
such a perspective, away from excessively structuralist views, the
observer, observation and observatum fuel the epistemic core act in its
interconnection rather then in their sharp separation. Science would
then always be defined in operational context, and not with a hypo-
static level of absolute objectivity anymore.
In this sense a new epistemology would be built, opposed to
the common sense one of the kosmotheoros, that is one of the «big
objects», in which the world stands, independent from observing
and awaiting for an observer to produce a neutral description of
it. The latter determination of epistemological attitude, even if it
was the enduring take on science until recently, is unable to
account for the complexity of all mechanisms enacted during
observation. The essay sketches a deck of ten objections to such a
model, which becomes also a programmatic statement of intent
for a new epistemology. Firstly, the paradigm of a non-relational
observation does not take into consideration the duplicity intrin-
sic in the observer figure, who is at the same time a living being, a
natural entity (similar to the observatum) and a human being de-
Book Reviews 179

fined in its ability to differentiate from its context namely through


the (self) reflective act of observation. Furthermore, observation
takes place in the mind but is enacted by and within a body,
which is in itself vehicle of additional information and variability
at the same time. Secondly, the act of observation does not exist
out of a context, which is not only the actual physical surround-
ings, but most especially (and long forgotten) cultural and histori-
cal ones. Finally the observatum never comes in total passivity, but
most of the times and especially in biological research appears in
action and inter-action with the observer. While the latter consid-
eration hence opens the necessity of analyzing a complex deck of
participation, communication, and recognition modes, under a
perspective where bio-philosophy immediately equals to bio-
ethics, the previous two show how relevant the opening of bio-
logical methodology could be to insights coming from other
disciplines such as psychology and neurosciences in order to
build a new epistemological framework.
«Frontiere della Biologia» is definitely a useful tool to all
scholarly levels: while it can be the perfect guide to new direc-
tions of research for both experienced scholars and beginners, it
does provide helpful insights in support of studies which already
focus on bio-philosophical essays, while also executing a peculiar
role of a boundary exploration. In fact, it brings many different
backgrounds to the table, calling all further efforts in research
towards a common ground, which will define, and orient all their
individual differences.

(Elena Tripaldi)

ANDREA BORGHINI, ELENA CASETTA, Filosofia della biologia, Ca-


rocci, Milano 2013, 307 pages, ISBN 978-88-430-6951-4.

The idea behind the newly-published volume Filosofia della bi-


ologia, by Andrea Borghini and Elena Casetta, is that the funda-
mental concepts used by science remain unfulfilled in the sphere
of scientific practice, which, from their viewpoint, implies the
necessity to resort to philosophy. The authors appear to imply the
above-mentioned position, as they take an interest in those no-
180 Book Reviews

tions determining the field of biology, the definition of which


seems to be still an open question.
The text presents itself as a freshly written, engaging intro-
duction to the philosophy of biology, full of examples and precise
explanations through which the authors lead the reader in the
gradual comprehension of the relevance and implications of the
main topics related to the life sciences.
Half a century ago a biologist’s reference to philosophy – e.g.
to Aristotelian naturalism – would still have been either proof of
mere erudition or a rhetorical expedient. On the other hand,
philosophy of science would exclude biology and its theories
from its field of enquiry, favouring physics. Instead, the idea
emerging from Borghini and Casetta’s text is that the dialogue
between the two disciplines – philosophy and biology – can be
fruitful for both sides. As a matter of fact, the two can first of all
find a meeting point in the philosophical roots from which biolo-
gy is originated. Secondly, contrasting the idea that often saw
philosophy and biology as antagonist to each other, one can also
meet a common horizon in which the efforts of both harmonise
under the guide of the famous delphic motto: as the authors
recall, philosophers are not the only ones to ask what man is.
On the one hand, the text underlines how the revolution in-
trinsic to biology never really lingers as such, as it can heavily
affect our lives by perturbing the idea we have of ourselves as
human beings. Simple examples of the subversion that can be
caused by biological theories could be Darwin’s theory (which
called into question the biological superiority of human beings) or
the discovery of DNA. This, however, creates a breathing space
for philosophy. Indeed, philosophy of biology finds its spot «in
the intersection between the facts biologists make us aware of
and the way in which those same facts end up influencing the
understanding that we have of ourselves» (p. 23). On the other
hand, as the authors highlight, there are different situations in
which biology is influenced by philosophy. First, as philosophy of
science, it can provide both important conceptual clarification and
an inter-theoretical integration. Second, philosophy assists biology.
One example could be the case of taxonomy, as «the contempo-
rary ontological and metaphysical inquiry» aiming to «categorise
Book Reviews 181

what there is in the world» and «an inquiry on the theories that are
presupposed (often implicitly) by biological classifications» (p. 25)
could provide a solid basis for an effective interaction between the
results obtained in the philosophical field and the work of the tax-
onomists. Thirdly, philosophy addresses biology. Again, an example
could be the case of biotechnologies, as biological research often
enough touches topics of deep public interest (i.e. GMOs, research
on stem cells or the debate on living will) or imposes decisions on
the allocation of research funding and therefore causes the philoso-
pher’s role to becomes crucial. In fact, it is no coincidence that
bioethics have a leading role in these matters. Moreover, philosophy
influences biology by first valuing those theories that have become
such an integral part of biological culture that they are no longer
seen as theories. Also, those very same theories, remaining untold
premises, could frustrate scientific research.
The book is divided into three main parts. The first is enti-
tled Processes and focuses on the processes of life and evolution
with a historical perspective. The authors put us in touch with
biology’s main topics: the discovery of DNA, Darwin’s theory of
evolution and Mendelian genetics.
What does it mean to be alive? This is the question ground-
ing the first chapter of the book. Tracing the boundaries of living
and nonliving implies many theoretical issues, which mainly
derive from the ambiguity of the terms in play. The idea is there-
fore to clarify the concept of life. The writers explain how nowa-
days we have discovered some chemical properties of the genetic
material that crucially inscribe the biological systems. In 1943
Erwin Schrödinger formulated a conjecture, claiming that the
aperiodic structure of the genetic material serves as division
between the crystals we find in the living systems and those that
compound the non-livings. This opened a new research trend
which led to the 1953 discovery of the double helix by Watson
and Crick and, later, to the unprecedented development of mo-
lecular biology. Ultimately, the aim of this chapter is to truly
understand the importance of these discoveries and their philo-
sophical consequences.
The second chapter is entirely dedicated to Darwin. «Nothing
makes sense in biology – claims Dobzhansky – unless you see it
182 Book Reviews

in the light of evolution»: the publishing of The Origin of Species


on the 24th of November 1859, constitutes one of the most
relevant events not just for biology, but also for human
knowledge in general. In this chapter Charles Darwin and his
works are contextualised with reference to the theoretical results
of the pre-evolutionary thought (from Carl Linnaeus’s taxono-
my to Thomas Robert Malthus’s studies on populations) and the
figures that made the birth of evolutionism possible (from Jean-
Baptiste Lamarck to Charles Lyell). From this starting point, the
authors conduct an efficient exposition of the theory of evolu-
tion synthesizing it in four simple fundamental ideas and show-
ing the philosophical and scientific richness intrinsic to the
evolutionary paradigm. If however the theory of evolution is
supported by a lot of empirical evidence (mostly provided by
The Origin itself) there are also some theoretical problems Dar-
win himself was already aware of.
One of the above-mentioned problems in the Darwinian
theory lays in the lack of a convincing explanation for hereditary
mechanisms. In fact, Mendelian genetics will overcome this
difficulty by explaining heredity according to quantitative princi-
ples and probabilistic laws. The third chapter is therefore dedicated
to the presentation of Gregor Mendel’s work in light of the
development of the evolution theory after Darwin. The authors
show how most of the early geneticists adopted an anti-
Darwinian perspective: the problem was to harmonize the continu-
ity of evolution by means of natural selection with the discrete
changes observable in genetic experiments. Such harmonization
was made possible by the scientific research undertaken in the
first decades of the 20th century, thanks to which many scientists
developed Mendel’s principles for the hereditary mechanism,
which also lead to the formulation of the so-called «New Synthe-
sis» in the late 40s. This theory understands Mendelian heredity as
an individual phenomenon and evolutionary selection as related
to populations. Distinguishing the two levels allows the conceiv-
ability of both theories, but also their complementarity.
Another aspect of Evolution after Darwin (which is also the ti-
tle of this third chapter) is that of the contemporary theory of
evolution. This is also the topic of the last part of the chapter,
Book Reviews 183

which distinguishes between naturalists and ultra-Darwinists. The


former understand natural selection as one – but not the only one
– of the means of modification of organisms: chance represents
another fundamental element. The latter, on the other hand,
affirm that evolution works in an algorithmic way, which allows it
to explain alone the evidence that we encounter in nature. Ultra-
Darwinists are however often methodological reductionists: for
example, for Dawkins (author of The Selfish Gene) evolutionary
phenomena can only be explained by considering genes as bearers
of a leading role in evolution. The bone of contention regards the
role of chance in evolutionary processes, both in the case of
hereditary mechanisms and for natural selection. The chapter
ends with a reference to two topics closely related to Darwinism:
creationism and the argument of intelligent design.
The third chapter and, with it, the first section of the book,
ends with a discussion on the developments of the post-
Darwinian theory of evolution. The second part is defined by the
writers themselves «more distinctly metaphysical» (p. 28). Its title,
Entities, recalls the topic: the authors are here considering those
entities generally admitted in the dominion of biology and prob-
lematizing the fundamental terms that normally guide both the
biological theory and practice. The first entity to be taken into
consideration in the fourth chapter is biodiversity. The word biodiver-
sity was coined during a national forum in Washington in 1986.
From that moment on, biodiversity and its safeguard have been a
very popular topic, exceeding the scientific field and spreading in
the public sphere. In cases like this, according to Borghini and
Casetta, philosophy can play an important role, first of all by
clarifying the meaning of the term. Secondly, biodiversity «is born
together with its crisis» (p. 98), namely it arose in the very mo-
ment in which biologists considered its safeguard necessary.
Consequently, biodiversity lays in the intersection of different
disciplines – ecology, epistemology, ethics, politics – among
which philosophy necessarily plays a leading role. In this chapter,
so as to highlight the ambiguity of the word, the authors first of
all analyse the nature of the concept biodiversity. They highlight
how its elusiveness derives from the vagueness of diversity as a
word itself, a concept both undefined from a formal perspective
184 Book Reviews

and intuitive, rooted in common sense. Moreover, preserving


diversity mainly involves knowing which diversities are to be pre-
served. As the tight bond connecting biodiversity and taxonomy
is now clear, the authors can present the fundamental ideas and
questions of both Linnaean and post-Darwinian taxonomies.
Starting from this analysis, the second part of the chapter takes
into consideration the different approaches that have been pro-
posed to describe biodiversity and tries to give an answer to two
fundamental questions: what to preserve? And, why preserve?
As mentioned before, taxonomy is a fundamental element in
biology and, among the entities it makes use of, an important role
is played by the species. From a historical perspective, species
appears to be a constant landmark for different epochs and
scientific paradigms from Aristotle to Linnaeus, from 15th and
16th century herbariums to Buffon. But how should we define
species? Is it just an expression or do species actually exist in the
world? What kind of entity are they? The fifth chapter tries to
address these questions, concerned with the so-called «problem
of the species». On the one hand, the effort is that of retracing
the history of the term, presenting the main concepts of species
that are still in use. On the other hand, the authors attempt to
confront the problems arising when trying to define the species,
and the ontological matters deriving from traditional realism. The
latter is effectively intrinsically monist and essentialist, features
that do not really harmonise with an evolutionary anti-fixity
understanding of the species. These problems can nevertheless be
reconsidered by other position like nominalism, pluralism or the
understanding of species as individuals (and, as such, subordinated
to the evolutionary process).
The key entity considered in the sixth chapter is the organism.
The reconstruction of the secular debate on what organisms are is
left aside to concentrate on the difficulties that, to date, could
inhibit any theory trying to explain the nature of organisms. The
writers, therefore, critically present some solutions offered by
contemporary authors. Asides from being fundamental in biolo-
gy, the notion of organism is an important element in our every-
day relation to the world. This is why, after outlining the tight
connection bonding the notions of life, organism and individual, the
Book Reviews 185

chapter focuses on our pre-scientific or phenomenal understand-


ing of organisms. This understanding is mainly anthropocentric
and proves to be insufficient when clarifying both the entities to
be found out of the ordinary, where our intuition and our lan-
guage are no longer solid guides, and in the most paradigmatic of
all organisms, namely the Homo sapiens. Given this insufficiency,
the last few pages of the chapter are dedicated to the exposition
of a different theory of organism. Moving from descriptive to
prescriptive metaphysics, the attempt is that of reconsidering our
ordinary understanding of organisms, by analysing the immunologi-
cal theory according to which the organism can be identified by the
presence of an immune system.
When confronting the problem regarding the nature of organ-
isms, the perspective adopted by the authors is that of reinserting it
in the wider debate concerned with biologic individuality: organisms
are biologic individuals among others (genes, cells, species, etc.). If,
therefore, the sixth chapter is dedicated to individual organisms, the
seventh chapter will be concerned with biologic individuals in general.
The authors start off by considering the notion of individuality.
They then follow a recent tendency adopted by biologic research
and try to retrace the theoretical path that has led from the enquiry
on organisms to that on biologic individuality. This path is two-
folded. First of all, it has a philosophical profile, which recalls both
Jack Wilson’s pluralist theory (which individuates six different types of
biologic individualities) and Robert A. Wilson’s theory (which under-
stands individuals as biologic agents). Secondly, it is related to the
theory of evolution and tries to identify the biologic individuals
with the unities of natural selection. The latter is also related to
Peter Godfrey-Smith’s theory of Darwinian individuals. We thus
reach the end of the section dedicated to entities.
The third (and last) part of this text, Implications, presents some
general topics related to biologic studies that are particularly
interesting from a philosophical perspective. Borghini and Casetta
argue that a debate between biology and philosophy is both
necessary and fruitful. This is also valid with regards to those
topics that exceed the scientific field and interest civil society and
public opinion. The eighth chapter addresses a particular aspect of
evolutionary theory, namely sex and sexual selection. The writers
186 Book Reviews

indicate six different meanings of the term sex, namely; sexuality,


process of mating, reproductive process, sexual character (male/female),
genus. The writers refer the terms to their semantic context and
make them interact with each other. This interaction results in the
following questions: what is the function of sex in the evolution-
ary process? Which criteria do we apply when we place an indi-
vidual in a specific sexual category? Is it truly that easy to divide
the human species in males and females? In the last part of the
chapter the distinction between sex as sexual character and as genus
is taken once again into account. As is known, this is a beloved
topic in philosophy.
When presenting the first chapter, we mentioned the revolu-
tion brought by Schrödinger’s ideas and by Waston-Crick’s dis-
covery of DNA. Thanks to the latter, biotechnologies were born in
the 70s. From that moment on, biotechnologies represent a field
of extraordinary scientific innovation and excite a lot of interest
both from the scientific perspective and in public opinion. This
has important philosophical implications and is the topic of the
ninth chapter. First, Borghini and Casetta retrace the historical
phases of the biotechnology revolution and analyse the results so
far obtained thanks to its research. They thereby linger on some
of the most important scientific endeavours, the most important
of which by far is the Human Genome Project. This project,
started in 1990 and ended in 2009, received a 3.8 billion dollar
loan and aimed to identify and map the human genome. The next
topic taken into account by the writers are the biopatents, which
are one example of the cases in which philosophy of biology can
play an important role and dialogue with politics. Next, they
analyse the development of agroindustrial and pharmaceutical
industries and thereby the birth of the scientist-businessman,
allowing the often ambiguous relation between economy and
science to emerge. Lastly, they take cloning into consideration.
The tenth chapter, Other Evolutions, confronts non-genetic evo-
lutions. These can be divided into three typologies: epigenetic
evolution, evolution of behaviour and symbolic evolution. The
first opens to a concept of heredity that is not solely related to the
genetic factor. Indeed, it engages those cases in which organisms
provided with the same genetic material display phenotipic and
Book Reviews 187

heritable diversities. The second is, on the other hand, related to


«a transmissible modification, produced by the social interactions
with other organisms usually belonging to the same species» (p.
243). Lastly, symbolic evolution is concerned with the human
ability to manipulate symbols from an evolutionary point of view:
the most relevant perspectives are Dawkin’s meme theory and the
90’s module theory, developed by some evolutionary psychologists.
Another form of non-genetic evolution is cultural evolution. Starting
from the analysis of these last two types of evolution – behav-
ioural and symbolic – the writers reinterpret one of the most
complex topics in philosophy of biology; namely, the relation
between nature and culture. The adopted perspective is clear: «the
culture of a human being is also his nature» (p. 258). Through the
examples provided by race and the evolution of altruism, the writ-
ers come to reflect upon the most appropriate methodology to be
applied when studying issues concerned with both nature and
culture. The assumed standpoint is that which avails itself of the
most recent biologic knowledge to make it interact with that
provided by human sciences.
The way in which philosophy and biology come to interact in
this text is exquisitely fecund. This is, first of all, because of the
constant conceptual enquiry aimed to clarify the terms in play
both from an ordinary and from a scientific perspective. The
latter is intended to avoid the misunderstandings that can emerge
from semantic ambiguities, both from the scientific and philo-
sophical viewpoints. We however believe that the truly original
element provided by Borghini-Casetta’s book lays in the deep
continuity established between philosophy and science. Indeed,
once we overcome an idea of philosophy of science as meta-
discourse on the methodological foundations of a particular
scientific discipline, the philosopher can have access to the scien-
tific practice and come to grips with the conceptual problems
related to the entities the scientist has to deal with during his re-
searches. As the authors say: «rather, what we are willing to un-
derline it that philosophy has something to say to and with biolo-
gy, and not only about biology» (p. 24).
What we have in front of us is therefore «a panoramic of the
philosophical foundations of biology», panoramic that, as men-
188 Book Reviews

tioned, is not historical, but rather conceptual. Indeed, apart from


the first section, where the speculative implications of scientific
topics are shown, the text aims to express biology’s main conceptu-
al cores. This is however combined with continuous and rigorous
references to authors and texts that orient the reader historically.
In conclusion, Philosophy of Biology is first of all an extremely useful
tool for whoever is willing to approach the discipline through a
nicely-flowing, up-to-date yet rigorous text. Last but not least, it is
also a noteworthy and valuable work for the recent philosophical
panorama.

(Daniele Bertoletti)
Finali:Layout 1 16-11-2009 10:24 Pagina 1

Pubblicazioni di Verifiche

Angelo Gemmi
La protologia nel pensiero di Gustavo Bontadini (esaurito)
Plutarco
Gli opuscoli contro gli stoici, 2 voll. (esaurito)
Storia del Trentino contemporaneo dall’annessione all’autonomia, 3 voll. (esaurito)
Giulio Lucchetta
Una fisica senza matematica (esaurito)
Franco Biasutti
Assolutezza e soggettività
L’idea di religione in Hegel
172 pagine - isbn 88-88286-10-1 (Pubblicazioni 5)

Franco Chiereghin
Dialettica dell’assoluto e ontologia della soggettività in Hegel
Dall’ideale giovanile alla «Fenomenologia dello spirito»
473 pagine - isbn 88-88286-11-x (Pubblicazioni 6)

Vincenzo Milanesi
Prassi e psiche
Etica e scienza dell’uomo nella cultura filosofica italiana del primo Novecento
213 pagine - isbn 88-88286-12-8 (Pubblicazioni 7)

Antonio Moretto
Hegel e la «matematica dell’infinito»
432 pagine - isbn 88-88286-13-6 (Pubblicazioni 8)

Franco Chiereghin
Essere e verità
Note a «Logik. Die Frage nach der Wahrheit» di Martin Heidegger
172 pagine - isbn 88-88286-14-4 (Pubblicazioni 9)

Adriana Cavarero
L’interpretazione hegeliana di Parmenide (esaurito)
Franco Biasutti
Ricerche sulla fortuna di Spinoza nel Settecento italiano (esaurito)
Francesca Menegoni
Finalità e destinazione morale nella «Critica del giudizio» di Kant
158 pagine - isbn 88-88286-16-0 (Pubblicazioni 12)

Antonio Moretto
Questioni di filosofia della matematica nella «Scienza della logica» di Hegel
«Die Lehre vom Sein» del 1831
107 pagine - isbn 88-88286-17-9 (Pubblicazioni 13)
Finali:Layout 1 16-11-2009 10:24 Pagina 2

Paolo Zecchinato
Giustificare la morale
230 pagine - isbn 88-88286-18-7 (Pubblicazioni 14)

Livia Bignami
Concetto e compito della filosofia in Hegel
221 pagine - isbn 88-88286-19-5 (Pubblicazioni 15)

Franco Biasutti
Prospettive su Spinoza (esaurito)
Franco Chiereghin
Il problema della libertà in Kant (esaurito)
Gabriele Tomasi
Identità razionale e moralità
Studio sulla «Fondazione della metafisica dei costumi» di Kant
166 pagine - isbn 88-88286-20-9 (Pubblicazioni 18)

Cristiana Bonelli Munegato


Johann Schultz e la prima recezione del criticismo kantiano
272 pagine - isbn 88-88286-21-7 (Pubblicazioni 19)

Francesca Menegoni
Soggetto e struttura dell’agire in Hegel
207 pagine - isbn 88-88286-00-4 (Pubblicazioni 20)

Gabriele Tomasi
Il «salvataggio» kantiano della bellezza
112 pagine - isbn 88-88286-01-2 (Pubblicazioni 21)

Luca Illetterati
Natura e ragione (esaurito)
Luca Illetterati
Figure del limite
Esperienze e forme della finitezza
115 pagine - isbn 88-88286-03-9 (Pubblicazioni 23)

Pietro Faggiotto
La metafisica kantiana della analogia
Ricerche e discussioni
204 pagine - isbn 88-88286-05-5 (Pubblicazioni 24)

Stefano Fuselli
Forme del sillogismo e modelli di razionalità in Hegel
Preliminari allo studio della concezione hegeliana della mediazione giudiziale
260 pagine - isbn 88-88286-06-3 (Pubblicazioni 25)

Paolo Giuspoli
Verso la «Scienza della logica»
Le lezioni di Hegel a Norimberga
xxii, 293 pagine - isbn 88-88286-07-1 (Pubblicazioni 26)
Finali:Layout 1 16-11-2009 10:24 Pagina 3

Alessandra Cover
Essere e negatività
Heidegger e la «Fenomenologia dello spirito» di Hegel
257 pagine - isbn 88-88286-08-x (Pubblicazioni 27)

Barbara De Mori
Cosa sono i diritti morali?
Un punto di vista analitico
179 pagine - isbn 88-88286-24-1 (Pubblicazioni 28)

Antonio-Maria Nunziante
Monade e contraddizione
L’interpretazione hegeliana di Leibniz
250 pagine - isbn 88-88286-09-8 (Pubblicazioni 29)

Giorgio Erle
La prospettiva di Hegel su tempo e natura
Ricerche e discussioni
xvi, 167 pagine - isbn 88-88286-26-8 (Pubblicazioni 30)

Giorgio Erle
Sul rapporto tra «ethos» e «physis» nella interpretazione hegeliana della filosofia greca
xviii, 114 pagine - isbn 88-88286-27-6 (Pubblicazioni 31)

Giorgio Corà
Verità ed entità affini
La verità tra scetticismo, analisi linguistica e speculazione
x, 321 pagine - isbn 88-88286-29-2 (Pubblicazioni 32)

Antonio-Maria Nunziante
Organismo come armonia
La genesi del concetto di organismo vivente in G.W. Leibniz
xi, 206 pagine - isbn 88-88286-30-6 (Pubblicazioni 33)

Nicoletta De Cian
Redenzione, colpa, salvezza
All’origine della filosofia di Schopenhauer
xx, 289 pagine - isbn 88-88286-31-4 (Pubblicazioni 34)

Mario Baggio
Port-Royal: dai modelli speculativi alla grammatica esplicativa (esaurito)
Lucia Procuranti
Il problema della costituzione della materia nella filosofia di Immanuel Kant
303 pagine - isbn 88-88286-33-0 (Pubblicazioni 36)

Monica Bassanese
Heidegger e Von Uexküll
Filosofia e biologia a confronto
xvi, 330 pagine - isbn 88-88286-34-9 (Pubblicazioni 37)

Frans Hemsterhuis e la cultura filosofica europea fra Settecento e Ottocento


A cura di L. Illetterati e A. Moretto
224 pagine - isbn 88-88286-35-7 (Pubblicazioni 38)
Antonio-M. Nunziante
Lo spirito naturalizzato
La stagione pre-analitica del naturalismo americano

Andrea Altobrando
Esperienza e Infinito: contributo per una
fenomenologia dell’idea di infinito a partire da Husserl

Barbara Santini
Soggetto e fondamento in Hölderlin tra
filosofia trascendentale e pensiero speculativo
Finali:Layout 1 16-11-2009 10:24 Pagina 5

Quaderni di Verifiche
Collana di testi filosofici diretta da Franco Chiereghin

Georg Wilhelm Friedrich Hegel


Scritti di filosofia della religione (esaurito)
Leo Lugarini - Manfred Riedel - Remo Bodei
Filosofia e società in Hegel
A cura di F. Chiereghin
172 pagine - isbn 88-88286-25-x (Quaderni 2)

Martin Heidegger
Hegel e i greci (esaurito)
Georg Wilhelm Friedrich Hegel
Logica e metafisica di Jena (1804-1805)
A cura di F. Chiereghin
Introduzione, traduzione e commento di F. Biasutti, L. Bignami, F. Chiereghin, A. Gaiarsa, M. Giacin,
F. Longato, F. Menegoni, A. Moretto, G. Perin Rossi
546 pagine - isbn 88-88286-22-5 (Quaderni 4)

Georg Wilhelm Friedrich Hegel


Enciclopedia delle scienze filosofiche in compendio (1817)
A cura di F. Chiereghin. Traduzione e commento di F. Biasutti, L. Bignami, F. Chiereghin, G.F. Frigo,
G. Granello, F. Menegoni, A. Moretto
279 pagine - isbn 88-88286-23-3 (Quaderni 5)

Filosofia e scienze filosofiche nell’«Enciclopedia» hegeliana del 1817


Contributi di F. Biasutti, L. Bignami, F. Chiereghin, P. Giuspoli, L. Illetterati, F. Menegoni, A. Moretto
608 pagine - isbn 88-88286-02-0 (Quaderni 6)

Georg Wilhelm Friedrich Hegel


Sul meccanismo, il chimismo, l’organismo e il conoscere
Introduzione, traduzione e commento di L. Illetterati
137 pagine - isbn 88-88286-04-7 (Quaderni 7)

Georg Wilhelm Friedrich Hegel


Logica e sistema delle scienze particolari (1810-11)
Introduzione, traduzione e note di P. Giuspoli
X, 248 pagine - isbn 88-88286-28-4 (Quaderni 8)

Georg Wilhelm Friedrich Hegel


Philosophische Enzyklopä die / Enciclopedia filosofica (1808-09)
Testo tedesco con traduzione a fronte, a cura di P. Giuspoli
X, 248 pagine - isbn 88-88286-37-3 (Quaderni 9)

Gottlob Ernst Schulze


Vorlesung über Metaphysik / Corso di Metafisica (1810-11)
Testo tedesco a cura di N. De Cian e J. Stollberg
Introduzione, traduzione e apparati di N. De Cian
X, 348 pagine - isbn 978-88-8828-641-9 (Quaderni 10)
Georg Wilhelm Friedrich Hegel
Scienza della Logica / Libro primo. L’essere (1812)
A cura di P. Giuspoli, Giovanni Castegnaro e Paolo Livieri
532 pagine - isbn 978-88-8828-642-6 (Quaderni 11)
Finito di stampare
nel mese di Luglio 2015
dalla Litocenter s.r.l. di Piazzola sul Brenta (Pd)

S-ar putea să vă placă și