Vous êtes sur la page 1sur 144

1

2
3
4
5
6
I. Introduction​ ​…………………………………………………………………………...9
II. Chapter​ ​One:​ ​The​ ​Logic​ ​of​ ​Inquiry​ ​………………………………………………….16
A. Inference​ ​and​ ​the​ ​Critique​ ​of​ ​Cartesian​ ​Intuitionism………………………...16
B. The Fixation of Belief-Inquiry, Scientific Method, and the Hypothesis of
reality………………………………………………………………………....27
C. The​ ​Real​ ​and​ ​the​ ​Critical​ ​Control​ ​Of​ ​Inference……………………………...40
III. Chapter​ ​Two:​ ​Semiotic…………………………………………………………….....47
A. Saussure​ ​and​ ​Signification…………………………………………………...48
B. The Origin of Peirce’s Semiotic in the Derivation of the
Categories………...50
C. The​ ​Triadic​ ​Theory​ ​of​ ​Signs​ ​and​ ​the​ ​Need​ ​for​ ​Interpretative​ ​Control…….....53
IV. Chapter​ ​Three:​ ​Pragmatism…………………………………………………………..62
A. Apel's​ ​Situation​ ​of​ ​Peirce​ ​in​ ​Kant’s​ ​Project………………………………….62
B. The​ ​Pragmatic​ ​Maxim………………………………………………………..64
C. Pragmatism​ ​and​ ​the​ ​Threat​ ​of​ ​Nominalism………………………………......67
V. Chapter​ ​Four:​ ​Normative​ ​Science……………………………………………………73
A. Peirce​ ​and​ ​Kant​ ​on​ ​Aesthetics………………………………………………..77
B. The​ ​Aesthetic​ ​Ideal​ ​in​ ​Critical​ ​Deliberation………………………………....83
C. The​ ​"New"​ ​Pragmatic​ ​Maxim……………………………………………......85
D. A​ ​Prolegomena​ ​for​ ​Metaphysics…………………………………………......87
VI. Chapter​ ​Five:​ ​Chaos​ ​Theory​ ​………………………………………………………...91
A. Definitions……………………………………………………………….…...93
B. Phase​ ​Space​ ​and​ ​the​ ​Mapping​ ​of​ ​Dynamical​ ​Systems…………………….....94
C. Strange​ ​Attractors………………………………………………………….....97
D. The​ ​Fractal​ ​Dimension…………………………………………………...…..99
E. Beyond Newtonian Determinism: Chaotic Systems and the Creative Vitality
of the
Cosmos……………………………………………………………….....102
F. Conclusion…………………………………………………………………..109
VII. Chapter Six: Metaphysics and Peirce's Evolutionary
Cosmology……......................111
A. Law, Spontaneity, and the Structure of the
Cosmos………………………...113
B. Evolution and the Categories - Tychism, Anancasten, and Agapism
……....117
C. Peirce​ ​and​ ​Chaos​ ​Theory……………………….……………………….......122
VIII. Conclusion…………………………………………………………………….....….133

7
Dissertation​ ​Abstract
STRANGE​ ​ATTRACTION;​ ​C.S.​ ​Peirce,​ ​Chaos​ ​Theory,​ ​and​ ​the​ ​Reclamation​ ​of​ ​Pragmatism
Darin​ ​McNabb
Richard​ ​Cobb-Stevens,​ ​advisor
The thought of C.S. Peirce was postmodern before is time. The pragmatic maxim and
his foundational work in semiotics offer potent challenges to traditional Cartesian
epistemology. The first three chapters of this dissertation examine Peirce's challenge to
Descartes and his transformation of Kant. However, Peirce tempered this success with
thoughts on normativity, metaphysics, and the Real. His thought, thcm, is poised between a
modernist metaphysics and a postmodern epistemology. As such my thesis concerns the
success with which he was able to do this. conclude that he was successful and my fifth
chapter​ ​adduces​ ​thc​ ​insights​ ​of​ ​chaos​ ​theory​ ​to​ ​help​ ​illustrate​ ​this.

Chapter One introduces his critique of Cartesian intuitionism. By conceiving of thought


as inferential and as carried out by a community of inquirers he reoriented the concepts of
judgment​ ​and​ ​truth​ ​to​ ​a​ ​publicly​ ​opcn​ ​future​ ​mediated​ ​by​ ​the​ ​praxis​ ​of​ ​scientific​ ​method.

Chapter Two examines the structure of inquiry in his conception of sign use. I compare
his conceptions of signs to that of Saussure's and examine the derivation of his categories as
the​ ​basis​ ​for​ ​his​ ​semiotic.

Chapter Three introduces the pragmatic maxim by which Peirce gave teeth to the course
of inquiry and by which meaning was severed from its Cartesian trappings. examine the way
in which a superficial understanding of the maxim leads to the nominalism which Peirce
despised.​ ​His​ ​attempt​ ​to​ ​amend​ ​it​ ​leads​ ​to​ ​the​ ​fourth​ ​chapter.

Here​ ​I​ ​discuss​ ​thc​ ​trio​ ​of​ ​his​ ​normative​ ​sciences​ ​--​ ​logic,​ ​ethics,​ ​and​ ​aesthetics,​ ​which
he​ ​conceives​ ​as​ ​indispensable​ ​in​ ​the​ ​approach​ ​inquiry

nakes​ ​to​ ​the​ ​Real.

Chapter Five concerns a discussion of chaos theory, whose conception of the strange
attractor clearly models the structure of the Real as given in Peirce's metaphysics. It also
provides​ ​a​ ​form​ ​of​ ​justification​ ​for​ ​Peirce's​ ​insights.

Chapter Six presents Peirce's conception of an evolving cosmos mediated by the


relation between law and spontaneity. The Real is seen to be neither absolute nor relative but
of a character which answers both the postmodern insights of his epistemology and the more
traditional​ ​notions​ ​of​ ​truth​ ​and​ ​reality.

8
Introduction

That C. S. Peirce is recognized as the father of Pragmatism is a commonplace, but what


passes under the name of Pragmatism in contemporary philosophy these days is far from the
vision Peirce had when he first formulated the term. In appreciating this ironic disparity, the
famous argument he had with William James over its meaning is instructive. The formulation
that Peirce gave it was in terms of a maxim, a methodological tool in the aid of inquiry. As it
is now famously known, "Consider what effects, which might conceivably have practical
bearings, we conceive the object of our conception to have. Then, our conception of these
effects is the whole of our conception of the object (5.402).”1 Though written in the
nineteenth century, it is one of the most revolutionary theses of twentieth century philosophy
for it completely broke open the mysterious regions of the Cartesian mind in which meaning
was held in some occult way to reside. It gave meaning a publicly observable quality and
accessibility hitherto occluded by competing epistemological theories grounded in what
Peirce​ ​argued​ ​to​ ​be​ ​the​ ​mistaken​ ​Cartesian​ ​assumptions​ ​of​ ​the​ ​nature​ ​of​ ​mind​ ​and​ ​reality.2

As mentioned, Peirce understood this redefinition of meaning as an aid in the course of


inquiry. He says of inquiry, in his paper ‘The Fixation of Belief’, that its sole object is the
settlement of belief (5.375). So it is not difficult to see, with such phrases as "practical
bearings", "effects", and "settlement of belief", how James came to understand meaning as
coined in his famous phrase "cash value." It was his interpretation of Peirce's maxim that the
upshot of a word's or theory's meaning was held in the role it played in settling practical
doubts, in enabling us to pursue our ends, whatever they may be. It is an interpretation in
some ways not substantially different from that held by Richard Rorty, who can be said to be
the standard-bearer for contemporary or neo- pragmatism. Like James, Rorty conceives
Peirce's insight in a practical or utilitarian fashion, allowing us to view language, words,
theories, the exchange of ideas, as tools enabling us to cope in the world. However he goes
further, in that pragmatism for him is not so much a maxim as it is a whole "climate of
opinion" about the usefulness of philosophy. It could be said in fact that he applies the maxim

1
​Charles S. Peirce. ​The Collected Papers of Charles S. Peirce. (Cambridge: Harvard University Press, 193-1935). The
conventional method used to cite Peirce's work employs a numbering system (as used above) where the number to the left of
the decimal refers to the volume number of ​The Collected Works​. and the number to the right to the paragraph therein.
Nearly​ ​all​ ​citations​ ​of​ ​Peirce​ ​here​ ​will​ ​be​ ​made​ ​in​ ​this​ ​way.
2
​These assumptions stem from the "container" theory of mind implicit in Descartes' positing of two kinds of substance:
thinking and extended, or mind and body, Generally, this epistemological view regards knowledge as being achieved when,
by whatever means, that which is outside the mind is brought inside. It is with this conception that contrasting views drawn
between​ ​Peirce​ ​and​ ​Cartesianism​ ​in​ ​general​ ​will​ ​refer.

9
to the notion of philosophy itself and finds that it's traditional concerns are illusory, having
their source in the unnecessary and misleading divisions erected by the reification of Truth.
As Rorty says, "A pragmatist theory of truth is not the sort of thing one should expect to have
a philosophically interesting theory about. For pragmatists, "truth" is just the name of a
property that all true statements share."3 The problem for him lies in the mad search for
foundations or invariant rules describing a paradigmatic language which mirrors reality. The
history of the tradition has only so far proven its fruitlessness and thus Rorty claims that the
pragmatist should seek to level these divisions and create a playing field in which all genres
of discourse - literary, scientific, philosophical, etc. -- are seen as equally meaningful in the
course​ ​of​ ​human​ ​expression.

Nothing could be further from Peirce's motivations as a thinker. If there can be found
one constant throughout all his writing it would probably be the insistent criticism he bore
down upon nominalism, and perhaps it could be said that none merit this criticism more than
Rorty himself. Although my discussion of Peirce will touch on his critique of nominalism,
my thesis does not directly concern a critique of Rorty's views. I mention him here as a foil to
Peirce's argument with James over the meaning of pragmatism and his re-appropriation of it
under the name pragmaticism.4 This move on Peirce's part involved metaphysical
commitments he found it necessary to hold in order to give pragmatism the effective scope he
intended for it in the course of inquiry. The question guiding my thesis thus regards, in a
broad way, the nature and plausibility of these commitments. I hope in some way to assuage
the irony alluded to in the beginning by showing that it is only within Peirce's metaphysical
defense of pragmaticism that his semiotic (which Rorty largely credits and champions Peirce
for) escapes the general relativizing thrust of contemporary theory. I will also argue (with
illustrations from the insights of chaos theory) that Peirce's metaphysics, born as a
consequence of his logical and categorial investigations, transcends the debilitating
restrictions​ ​inherent​ ​in​ ​a​ ​traditional​ ​metaphysics​ ​based​ ​on​ ​the​ ​Cartesian​ ​theory​ ​of​ ​knowledge.

I will begin by briefly describing Peirce's critique of Cartesianism, out of which will
arise his judgment that cognition is inferential in nature and operates, as does language-use,
in a triadic relationship of signification characterized by sign, object, and interpretant. This is

3
​Richard Rorty, "Pragmatism and Philosophy." In ​After Philosophy: End or Transformation​. Kenneth Baynes, James
Bohman,​ ​and​ ​Thomas​ ​McCarthy,​ ​eds.​ ​(Cambridge:​ ​The​ ​MIT​ ​Press,​ ​1987).​ ​p.​ ​26.
4
​ ​I​ ​will​ ​discuss​ ​this​ ​argument​ ​in​ ​my​ ​discussion​ ​of​ ​Peirce's​ ​pragmatism​ ​in​ ​Chapter​ ​Three.

10
the bare structure of his epistemology which is fleshed out in terms of his logic of inquiry. I
will elucidate the logic of inquiry by describing (according to Karl-Otto Apel’s reading)
Peirce's​ ​relationship​ ​to​ ​and​ ​transformation​ ​of​ ​Kant's​ ​transcendental​ ​framework.

In place of the unifying work of the synthetic unity of apperception I will show Peirce
substituting his semiotically based logic of inquiry. The logic of inquiry is based on the three
logical forms of argument: deduction, induction, and hypothesis, which lead Peirce to a
reformulation of the concept of meaning, given in his pragmatic maxim (stated above).
Meaning is no longer, for Peirce, the result of a Kantian unification of a manifold under a
concept, but is expressed in terms of a counterfactual conditional, or more colloquially, as the
formation of a habit or rule of action in regard to the conditional expectations we have about
how the object of our conception will behave. The habit is formed as a result of the process of
forming a hypothesis, deducing predictions from it, and inductively testing those predictions
against experience. Where Kant, in order to secure the normativity of the constitution of
meaning, had to perform a transcendental deduction of the validity of the categories, Peirce
must​ ​deduce​ ​the​ ​objective​ ​validity​ ​of​ ​synthetic​ ​inference.

The basis for this validity arises as a consequence of his redefinition of the real
necessitated by his critique of meaning. Kant's epistemological framework forced him to
define the real as ultimately unknowable, noumenal. But Peirce's pragmatic, semiotic
transformation of Kant's epistemology makes the notion of a thing in-itself nonsensical, for
the real that we reach through cognition is not so much "represented" as "embodied",
consisting, for Peirce, in the establishment of habits of action. For Peirce, the Real is only the
object of the final opinion to which sufficient investigation would lead. So the validity of
synthetic inference is given, in the long run, as the outcome of sufficient investigation by a
community​ ​of​ ​inquirers.

This argument is the lynch-pin upon which the bulk of his philosophy turns. As he said,
"But antecedently to this to Kant's question how synthetic judgments a priori are possible
comes the question how synthetic judgments in general, and still more generally, how
synthetic reasoning is possible at all . . . That is the lock upon the door of philosophy.
(5.348)"

11
What is too facilely assumed about Peirce's deduction of the validity of inference in the
long run is that it implies that a state in which all questions have been answered will actually
be achieved. This, in fact, is Rorty's main criticism of Peirce's "metaphysical pragmatism."
However, if this were Peirce's position he would not differ in any significant way from a
Hegelianism. Rather, the Real, understood in terms of the final opinion, is something that lies
in the infinite future and operates for Peirce as a regulative ideal, a methodological
assumption for the process of inquiry. The real for Peirce is indeed the generality and
lawfulness of nature, but where he differs from Hegel is in his insistence on the real operative
qualities of spontaneity and brute facticity in the universe, which do not become aufgehoben
in some Hegelian fashion.5 For Peirce, the lawful regularity of the universe is not something
already determinate which we discover through the process of inquiry, but is rather the
function of an evolutionary process. As he says, "The only possible way of accounting for the
laws of nature and for uniformity in general is to suppose them the results of evolution. This
supposes them not to be absolute, not to be obeyed precisely. It makes an element of
indeterminacy, spontaneity, or absolute chance in nature (6.13)." What makes some sort of
infinite concretization of the real impossible is the "tychastic" evolution of the universe.
Tychism, as Peirce calls it, is the evolutionary expression of the universe understood in terms
of the real events of fortuitous, spontaneous deviation from law. The conception of the
universe​ ​which​ ​evolves​ ​from​ ​this​ ​is​ ​one​ ​which​ ​is​ ​dynamic​ ​and​ ​evolving.

It is in connection with this view that I want to adduce the insights of chaos theory. I do
so in view of what I'm trying to reconcile in Peirce, namely the apparently conflicting
commitments of his early pragmatism and of his later pragmaticism, or in short, of the
modern and postmodern strains in his thought. I see in chaos theory a model for a similar
kind of reconciliation: one between a Newtonian vision of the universe as determinate and
thoroughly rational, and a contemporary one in which the lawful predictability of the world
we experience is reduced to the fortuitous and whimsical manifestations of a fundamental
chaos.6 Chaos theorists claim that the world is characterized not by either one or the other of

5
​Richard Rorty, "Pragmatism and Philosophy." In After Philosophy: End or Transformation. Kenneth Baynes, James
Bohman, and Thomas McCarthy, eds. (Cambridge: The MIT Press, 1987). p. 26. *I will discuss this argument in my
discussion​ ​of​ ​Peirce's​ ​pragmatism​ ​in​ ​Chapter​ ​Three.

6
​By this distinction I do not mean to say that contemporary scientists carry out research guided by the hypothesis of
discovering nothing but chaos, but I refer rather to the tendency of some contemporary cultural and critical theorists to refer

12
these visions in any determinate way but by both in a sort of dialectical relationship, as two
sides of the same coin. What chaos theorists study and try to account for is the rise of erratic,
unpredictable behavior in what appear to be otherwise progressively linear dynamic systems.7
As I will show, chaotic behavior which arises (whether in the torrential flow of a river, the
fluctuations of the stock market, or the sudden erratic beating of the heart) is best understood
not as fundamentally chaotic and unpredictable, which is how a determinist/reductive
framework would interpret it, but as indicative of a complex, evolving whole seen in a
qualitative, global sense. Chaos theorists characterize the variables in non-linear dynamic
systems as locally unpredictable but globally stable, much like the weather system. And it is
precisely this integrative, evolving sense of the world that I see operative in Peirce's
metaphysics. The various elements of his philosophy commit him to a position which is
neither completely determinist nor completely relativist. It is here where his early pragmatism
and later metaphysics seem to collide and call out for reconciliation, and it is in this
connection​ ​that​ ​I​ ​see​ ​as​ ​illuminating​ ​and​ ​helpful​ ​a​ ​discussion​ ​of​ ​chaos​ ​theory.

I hope to serve two interests in this discussion of chaos theory. First, as simply
providing a model to help illustrate, in concrete terms, the abstractly formulated framework
of Peirce's metaphysics. Secondly, in some real sense, to provide justification for his theory.
With regard to this second aim, it goes without saying that adducing the results of empirical
research to support a philosophical theory is traditionally risky at best, and at worst,
downright methodologically irresponsible, especially given Peirce's fundamental concern for
the normative role of logic at the basis of scientific inquiry. However, the role I see chaos
theory playing in my thesis is not a foundational one. It is given rather in the very spirit of
Peirce's epistemological program of inquiry -- the positing of an hypothesis, the deducing of
expected effects from the hypothesis, and the inductive testing of it in the course of our
experience. If Peirce's metaphysical conjectures are true then it is only reasonable that
evidence​ ​for​ ​them​ ​accrue​ ​in​ ​the​ ​course​ ​of​ ​scientific​ ​inquiry.​ ​As​ ​Peirce​ ​says,

to the chaotic behavior of dynamical systems as exemplary of the decenteredness and indeterminacy which they assert in
their​ ​theories.
7
​The object of study for chaos theory has its philosophical counterpart in the dislocation and unpredictability of meaning
and truth in contemporary theory. Philosophy in the twentieth century has torn asunder the “easy” metaphysics of the
tradition, revealing the dissonance and contingency underlying apparently rational foundations. I believe the insights of
chaos theory can helpfully illuminate the way in which Peirce’s metaphysical speculations address this contemporary
dilemma.

13
Philosophy ought to imitate the successful sciences in its methods, so far as to proceed
only from tangible premises which can be subjected to careful scrutiny, and to trust
rather to the multitude and variety of its arguments than to the conclusiveness of any
one. Its reasoning should not form a chain which is no stronger than its weakest link, but
a cable whose fibers may be ever so slender, provided they are sufficiently numerous
and​ ​intimately​ ​connected​ ​(5.265).

I offer chaos theory not as a Cartesian-like indubitable basis for justification but as one
"fiber" aiding in the growth of knowledge. The word "justification" may be too strong
though. In the terms in which I will describe chaos theory, I will intend for it not so much to
vindicate Peirce as to offer a warrant for giving serious consideration to his metaphyscial
speculations, for accepting them, provisionally, as admissably worthwhile hypotheses for the
work​ ​of​ ​inquiry.

In concluding my thesis I hope to show that Peirce's pragmatic reorientation of meaning


understood in the context of a metaphysics of evolution avoids the horns of the contemporary
relativists determinist dilemma mired in the legacy of Cartesianism. Understood in the
context of an evolving, pansemiotic universe, truth must lose the reified nature of it given in
the tradition, however it is not lost, just re-understood in terms of its embodiment in the
practices of a critical community of inquirers. The sort of truth or reality given in these
practices is, though always evolving, of a fundamentally stable nature which transcends the
relativizing​ ​distinction​ ​between​ ​dissemination​ ​and​ ​metaphysics.

As will be seen in the course of examining Peirce's logic of inquiry, one of the
presuppositions he makes regarding the activity of inquiry is that it proceeds according to a
guiding principle. Movement from a state of doubt or uncertainty to one of belief is
accomplished when we employ axioms or hypotheses about the world in constraining the
conclusions we arrive at in the formation of belief. Using Peirce's example, the guiding
principle, "that what is true of one piece of copper is true of another" is an hypothesis which
mediates the inference that if we "observe that a rotating disk of copper quickly comes to rest
when placed between the poles of a magnet," this "will happen with every piece of copper"
(5.367). The principle guides us in coming to a judgment in which the irritation of doubt has
been​ ​appeased.

14
It would seem, thus, remiss to begin an inquiry into Peirce's thought without
establishing some sort of guiding principle to aid in constraining possible interpretations.
What is broadly at issue in my thesis is the need to provide a consistent account of the
interaction of Peirce's pragmatism with the evolution of his metaphysical conjectures in the
last few decades of his life, to make sense of the metaphysical tincture enveloping what he
came to call pragmaticism. For Peirce this renaming was a decided move away from the
nominalism and subjectivism of James and reflects the life-long influence of the lessons he
learned from Kant, namely, of the need for a deeply rooted, encompassing architectonic
informed at its base by a sound logic.8 Though he was critical of Kant in many ways his
methodological temperament left a mark on Peirce which is evidenced throughout the broad
outlines of his thought to the very end. Thus an appropriate heuristic in the context of my
thesis would seem to be one which reflects this influence of Kant, which would lend
explanatory power to the transcendental concerns evident in the evolving refinement of his
pragmatism.

This particular thesis, that there's a largely Kantian cast to what Peirce is trying to
accomplish, has already been given forceful defense by Karl-Otto Apel.9 He argues that
although Peirce is revolutionary in many of his central theses about the nature of philosophy,
he remains largely loyal to the broad outlines of Kant's transcendental project. One of my
interests here, in terms of accounting for Peirce's pragmaticism, is to provide a way of
appreciating Peirce's fidelity to transcendental concerns while at the same time articulating
how his novel, semiotic approach to mind and reality furnishes a metaphysics which escapes
the foundationalist dilemma that Kant ultimately foundered upon. Using Apel's thesis as a
guide I hope to give context to the development of Peirce's epistemological insights, showing
how he diverges from Kant in many important ways, and showing as well that, in spite of the
transfigured context of transcendental questions brought about by his pragmaticism, he
thereby provides a fruitful ground for a metaphysics which eludes the subtle and disquieting
options​ ​for​ ​philosophy​ ​provided​ ​for​ ​in​ ​Rorty's​ ​version​ ​of​ ​pragmatism.

8
​For a criticism of the view that William James holds an exclusively nominalist position, see James M. Edie. ​William James
and​ ​Phenomenology​ ​(Bloomington:​ ​Indiana​ ​University​ ​Press,​ ​1987).

9
​ ​Karl-Otto​ ​Apel,​ ​Charles​ ​S.​ ​Peirce:​​ ​From​ ​Pragmatism​ ​to​ ​Pragmaticism​ ​(Amherst:​ ​University​ ​of​ ​Massachusetts​ ​Press,​ ​1981).

15
Chapter​ ​One

The​ ​Logic​ ​of​ ​Inquiry

In addition to Kant, one of the most pervading threads of influence to be found in Peirce
is his scientific training. Unlike his contemporaries James and Dewey, he was a practicing
scientist, having worked for the Harvard Observatory and U.S. Coastal and Geodetic Survey
for a number of years, and having contributed a number of significant advancements in
astronomical measurement, gravitometry, and pendulum research.10 To the great boon of
philosophy his interest was not purely scientific but philosophical as well. Early on, the
temperament forged by his scientific work was manifest where, in a passage from an
unpublished manuscript, he sees himself being among "those few fellow-students of
philosophy, who deplore the present state of that study, and who are intent upon rescuing it
therefrom and bringing it to a condition like that of the natural sciences" (5.413). To
contemporary ears such a sentiment lies at the very heart of the tribulations of modern
philosophy. Starting with Descartes' epistemological premises this endeavor has continually
foundered upon the incommensurability between the veiled mechanisms of the cogito and the
external world. What Peirce brings to this problem is a thoroughgoing experience of scientific
practice and a revolutionary epistemological framework to go along with it. Articulating the
elements of this epistemology requires an understanding of the logic underlying it, and the
critique​ ​of​ ​Cartesianism​ ​which​ ​made​ ​it​ ​possible.​ ​It​ ​is​ ​to​ ​this​ ​which​ ​I​ ​now​ ​turn.

Inference​ ​and​ ​the​ ​Critique​ ​of​ ​Cartesian​ ​Intuitionism

Peirce's critique of Descartes is framed in the larger context of an excursus on the


nature and foundational role of logic in inquiry, given most forcefully in a series of three
papers: "Questions Concerning Certain Faculties claimed for Man"; "Some Consequences of
Four Incapacities"; "The Fixation of Belief". Written before the age of thirty these essays
reflect an initial philosophical interest weaned on his early and deep reading of Kant's
Critique of Pure Reason. Though he greatly admired Kant's genius and insight, Peirce found
himself in the end faced with a metaphysics which seemed implausible, tainted with special
pleading, and contrary to his deepest instincts as a scientific investigator. He found the

10
​ ​See​ ​Brent,​ ​Joseph,​ ​Charles​ ​Sanders​ ​Peirce:​ ​A​ ​Life​,​ ​(Bloomington:​ ​indiana​ ​University​ ​Press:​ ​1993).

16
categorical forms of judgment postulated by Kant to be incomplete and reflective of what he
deemed to be an inadequate logic at their basis. One of the greatest lessons he would take
from Kant is the normative role of logic grounding any metaphysics, and it was from here,
from Kant's logic, that his disagreements stemmed. It would be his aim in a later paper11 to
revise Kant's categories, but in this important first series he concerns himself with a rigorous
weeding out of the faulty logical assumptions which spoiled Kant's efforts, assumptions Kant
unwittingly​ ​borrowed​ ​from​ ​Descartes.

Peirce is concerned in these papers to defend the autonomy of logic against the
rendering of it in terms of either of the two horns of the Cartesian epistemological dilemma as
he sees it. The dilemma is formed when the attempt to account for knowledge accepts what
can be called the 'container' theory of mind or cognition, the notion that knowledge is
achieved when we are able to represent the essence of something external to the mind, inside
the mind itself. Descartes accounted for the success of this ability, and guarded against its
failure, by appeal to the power of intuition, the direct, apprehension of something external to
the mind. For Descartes an intuition is a cognition (whether judgment or sensation) the
content of which reflects not the influence of prior cognitions which produced it (which he
distinguished as inference) but the direct action of the thing represented as it is in itself. The
horns of the dilemma rear themselves at this point in the recourse Descartes is compelled to
make in ensuring that our intuitions do not deceive us but actually represent reality as it is.
The horn he chose was to posit, and then attempt to prove, the existence of a benevolent God,
the inadequacy of which has been borne out by the persistent appearance of the other horn
throughout​ ​the​ ​history​ ​of​ ​philosophy​ ​since.

For Peirce the other horn is psychologism, or as he understood it in the larger class to
which it belongs, nominalism, an explanatory tendency he traces from Ockham, through
Hobbes, Berkeley, Locke, Hume, and Mill, up through and including his colleague and friend
William James. Still operative here is the basic epistemological model of "getting what's out
there inside", however Descartes' inarticulate account of intuition is refined in a technical
empiricism and the logic which replaces God's benevolent assurances becomes a
psychological account of how cognition forms judgments. For Peirce, however, this makes
logic dependent on that which it should ground, and leads any subsequent metaphysical view

11
​"On a New List of Categories," in ​Proceedings of the American Academy of Arts and Sciences. Vol. 7 (May 1867), pp.
287-98.

17
of reality to the nominalism which he despised. Nominalism is the view of reality which
holds that the "real" is constituted by particular, existent objects, knowledge of which is
somehow accomplished by intuiting them through the senses, making of general terms, such
as​ ​"hardness,"​ ​nothing​ ​more​ ​than​ ​mere​ ​conveniences​ ​in​ ​the​ ​classification​ ​of​ ​experience.

This view of reality has a lot of commonsense power behind it but is at a loss, in Peirce's
mind, to explain the "laws" which give meat to the successful course of scientific inquiry.
The role of cognition operating on a model which sees reality as nothing more than the
efficient cause of our sensations becomes wholly problematic, leading to a Kantian world of
things in themselves with no guarantee that our reasoning results in any true propositions
about that reality. The role of logic falls then to the Inere description of how we happen to
employ the use of our reason, severing the synthetic growth of scientific knowledge from any
basis or foundation by which it can be fruitfully explained. As mentioned above, explaining
this growth is the "lock upon the door of philosophy." Kant, the psychologists, and finally
Descartes, to whom the lineage is indebted, failed in this. It will be his account of inference
and​ ​more​ ​broadly​ ​of​ ​his​ ​semiotic​ ​by​ ​which​ ​Peirce​ ​hopes​ ​to​ ​succeed.

Peirce's account of the inferential nature of thought is constructed in the rubble of his
critique of Descartes. In "Questions Concerning Certain Faculties Claimed For Man" he asks
a series of seven questions concerning our assumptions about the nature of cognition and self
knowledge, and through the doubts raised in his answers begins to thread an epistemological
picture​ ​which​ ​will​ ​make​ ​way​ ​for​ ​his​ ​logic​ ​of​ ​inquiry.

He asks first if "whether by the simple contemplation of a cognition . . . we are enabled


rightly to judge whether that cognition has been determined by a previous cognition or
whether it refers immediately to its object [intuition].12 "Peirce replies that we have no
evidence of such a faculty other than that we feel to have it, an observation which he sees
leading to a slippery and most precipitous slope. For of the judgment that a cognition is
indeed a direct intuition it may be asked of that judgment, which is itself a cognition, whether
it too is intuitive in nature or rather determined by a previous cognition. Exercising such a
faculty,​ ​Peirce​ ​says,​ ​depends​ ​on​ ​"presupposing​ ​the​ ​very​ ​matter​ ​testified​ ​to."

It is not his concern here to deny either the existence of nor the possible role intuition
may play in cognition but merely to cast doubt on our ability to distinguish and identify them.

12
​ ​Philip​ ​P.​ ​Wiener,​ ​ed.,​ ​Charles​ ​S.​ ​Peirce:​ ​Selected​ ​Writings​,​ ​(New​ ​York:​ ​Dover​ ​Publications,​ ​1958),​ ​p.​ ​18.

18
Reliance on this supposed faculty must end as the role of academic and ecclesial authority did
in​ ​the​ ​Middle​ ​Ages.

Supposing however that our sense faculties operate in an intuitive fashion in furnishing
the objects of the world to cognition, Peirce goes further in calling into question even this
assumption. He adduces a series of observations which suggest that objects of sensation are
inferred rather than immediately intuited. "Does the reader know of the blind spot on the
retina?" he asks. By directing the reader to perform the simple operation by which every child
is astonished to learn of the blind spot, he renews, in a different context, an astonishment
which leads us to reconsider the nature of sense perception. Because the optic nerve is stuck
in the middle of each retina, disrupting the even field of perception, the brain must make
inferences from the data of surrounding receptors so as to create the continuous field of
vision we normally experience. The apparently unified field of vision which seems given as
whole​ ​in​ ​the​ ​act​ ​of​ ​seeing​ ​is​ ​really​ ​just​ ​a​ ​patchwork.

Peirce makes similar observations with regard to distinguishing different tones, and
different textures of cloth. We cannot make this comparison immediately but, as in the case
of different textured cloths, are required to run our hands over it, comparing the sensation of
one​ ​instant​ ​with​ ​that​ ​of​ ​another​ ​before​ ​a​ ​sensation​ ​of​ ​the​ ​texture​ ​as​ ​such​ ​is​ ​possible.

In citing these observations Peirce reminds us that a theory is proposed in order to


explain the facts. With regard to the facts he has cited, he says of them they are all "most
readily explained on the supposition that we have no intuitive faculty of distinguishing
intuitive from mediate cognitions." The theory of intuition is thus superfluous in this sense,
extra baggage which lends more confusion than clarification to the cognitive process. And
though the claims and observations he makes in the course of this paper may seem striking
and cogent, it may be objected that they are nowhere near conclusive. But this is not his
intent.​ ​His​ ​aim​ ​is​ ​not​ ​so​ ​much​ ​one​ ​of​ ​refutation​ ​as​ ​it​ ​is​ ​suggestion.

Having called into question the confidence we have about our intuitive capacities, he
proceeds, in the second question of "Questions Concerning Certain Faculties Claimed For
Man", to draw out the consequences of this in assumptions we have about self knowledge.
His question concerns whether or not we have an intuitive self-consciousness, and it is
appropriately asked here, for lying at the basis of Descartes' whole philosophy is the cogito,
the certainty of self which through immediate intuition founds the certainty of all knowledge

19
to follow. It is in doing away with this notion of intuition that Peirce will be able to give a
much​ ​clearer​ ​understanding​ ​of​ ​the​ ​cognitive​ ​process.

Again, as in the first question, he does not deny the existence of intuition but asks
merely whether self-consciousness can be explained by the action of known faculties or
whether it is necessary to suppose an indemonstrable cause such as intuition. In addressing
this he adduces observations about very young children. He notices that in them there is no
sense of self-consciousness comparable to what we feel as adults. He describes the milieu of
their experience as a complex of sensation and desire: the hearing of a bell, the taste of food,
the movement of objects that come in contact with the child's body. The body becomes an
important fact in the child's experience because it emerges as the locus of change consequent
upon desire. The child then begins to associate and connect with this change vocal sounds
which becomes the basis of the formation of language. In desiring something the child finds
that the confluence of certain sounds and bodily movements bring about the satisfaction of
the desire, but Peirce contends that there is no sense of a self as such bringing these
conditions about. It is only when desire is thwarted that self-consciousness comes to be
formed.​ ​In​ ​a​ ​famous​ ​passage​ ​from​ ​this​ ​paper​ ​Peirce​ ​describes​ ​a​ ​child​ ​touching​ ​a​ ​hot​ ​stove,

A child hears it said that the stove is hot. But it is not, he says; and, indeed, the
child's body is not touching it, and only what that touches is hot or cold. But he
touches it, and finds the testimony confirmed in a striking way. Thus he becomes
aware of ignorance, and it is necessary to suppose a self in which this ignorance can
inhere.​ ​So​ ​testimony​ ​gives​ ​the​ ​first​ ​dawning​ ​of​ ​self-consciousness​ ​(5.233).

According to Peirce a child comes to form an image of itself as self when it experiences error
and frustration in the world of its expectations. The self is not intuited, but rather grows in a
series of experiences in which the child must make sense of thwarted expectation. The child
recognizes a relationship between an unexpected phenomenon and that which expected it,
and​ ​in​ ​so​ ​doing​ ​supposes​ ​or​ ​infers​ ​a​ ​self​ ​as​ ​an​ ​hypothesis​ ​which​ ​explains​ ​the​ ​error.

In these first two questions we see Peirce forming the backbone of a new epistemology,
the most salient feature of which thus far is something not so much constitutive as it is
methodological -- that is, the paring and gleaning practices of the scientific method. The
measure by which he judges Cartesian claims is the methodological axiom that theories must
have explanatory power, and that it is those theories which most efficiently and observably

20
explain a phenomenon which must be accepted and pursued. The phenomenon here is
cognition, and the competing theories, intuition and inference. Though in the end Peirce will
have nothing to say about intuition as such, he will say that as a theory to explain cognition it
is​ ​simply​ ​not​ ​viable.

In the third and fourth questions of "Questions Concerning Certain Faculties Claimed
For Man" Peirce further pursues problems of introspection and the ability to distinguish
subjective from non subjective elements of cognition, but as they don't really contribute to the
pertinent thrust of his critique I will move to the fifth question. Here he intensifies his focus
and asks a question with broad implications for his critique, "Whether we can think without
signs." An affirmative answer to this would mean that though we may actually intuit some
things, anything held in cognitive awareness would be born of the result of some inference,
the interpretation of some sign. Following on the heels of the answer to his first question he
responds​ ​in​ ​the​ ​negative.​ ​Thinking​ ​must​ ​find​ ​its​ ​currency​ ​in​ ​sign​ ​activity.​ ​As​ ​he​ ​says,

... we have seen that only by external facts can thought be known at all. The only
thought, then, which can possibly be cognized is thought in signs. But thought
which cannot be cognized does not exist. All thought, therefore, must necessarily be
in​ ​signs.13

Only by external facts can thought be known at all. Are we able to discern whether a
cognition has an intuitive origin or not? No. Only by appeal to the testimony of external fact
does awareness of thought, and thus its cognitive potency, come into being. For Peirce,
thought is an organic phenomenon, analogous to the many notes of a musical composition.
The relation of the notes to each other over a given expanse of time creates the experience we
call music. As he observed in the first question, the aural impulses which go to create the
hearing of a tone cannot be had individually but must work in concert for the sensation of the
tone to be possible. So it is with a musical score. Only the relation of the many notes to one
another creates a distinctive musical phrase. And so it is also with thought. If thought were
absolutely isolated and not born of some relation to another thought, it could not persist in the
mind;​ ​it​ ​could​ ​not​ ​function​ ​cognitively.

Peirce continues even more explicitly. Not only is thinking without signs impossible,

13
​ ​Wiener,​ ​p.​ ​34

21
but every thought itself is a sign. A detailed examination of his semiology will come in
Chapter Two, but it is now sufficient to note that the potency of an individual thought in the
cognitive process depends on its ability to address other thoughts, at other times, thus acting
"significantly"​ ​and​ ​sustaining​ ​the​ ​cognitive​ ​flow​ ​that​ ​we​ ​call​ ​thinking.

The final question I will address in Peirce's paper is the sixth, which is his rejoinder to
the Kantian notion of the thing-in-itself. He asks, "Whether a sign can have any meaning, if
by its definition it is the sign of something absolutely incognizable." This notion is a direct
implication of the Cartesian epistemological assumptions that find themselves threaded
throughout Kant's work, and it is important for Peirce to address this problem for Kant
represents​ ​the​ ​horizon​ ​within​ ​the​ ​scope​ ​of​ ​which​ ​modern​ ​nominalism​ ​finds​ ​its​ ​definition.

Remember, it is the nominalist reduction of the realin of concepts to utilitarian tools,


and thus the restriction of reality to the objects of sense perception that Peirce finds
nonsensical in trying to understand cognition and the growth of knowledge. He intends to
break apart this structure in a way which will give the nature of thought, conception and
cognition a more pervasive and organic role in the inception and growth of knowledge. Part
of​ ​his​ ​answer​ ​to​ ​this​ ​question​ ​includes​ ​the​ ​following​ ​striking​ ​remark,

Over against any cognition, there is an unknown but knowable reality; but over
against all possible cognition, there is only the self-contradictory. In short,
cognizability (in its widest sense) and being are not merely metaphysically the same,
but​ ​are​ ​synonymous​ ​terms​ ​(W.​ ​2.208).14

Clearly no statement could more fully betray Peirce's opposition to Descartes and the
nominalists than one which equates cognizability with reality, with being. And perhaps it is
unwise to introduce this statement here, for a full understanding of it will require the
explication of his categories, however a few things can be noted. His question is whether or
not a sign can have meaning if it stands for something that is absolutely incognizable.
Something incognizable would be something wholly unrelated to thought. If a thing is not
related to thought then it cannot be conceived. But it is the very fact that we are considering a
sign of something not cognizable which is what Peirce finds contradictory. If it is conceived,
then​ ​it​ ​must​ ​be​ ​related​ ​to​ ​thought.

14
​Max Fisch, ed., Writings of Charles Sanders Peirce: A Chronological Edition, ​(Bloomington: Indiana University Press,
1982-1986),​ ​Wols.​ ​I-IV.

22
The argument here may sound more like a semantic quibble than anything else, and
indeed there is an element of that because the two sides are working with two different
vocabularies, but it serves as an occasion to glimpse the radical epistemological shift Peirce
has in mind. Perhaps Kant would grant Peirce that we do form conceptions of the
incognizable, however he would insist on our ultimately being unable to know it. Peirce's
response would be that the real, the knowable, is not the static object of perception but the
complex of relationships in which that object partakes in the world and which thus give scope
to its nature, to its knowability. The result of cognition, knowledge, does not mirror the
outside world, but rather embodies the manner in which the "things" of the world relate to
one another. It is in this sense that he can say that "cognizability (in its widest sense) and
being are not merely metaphysically the same, but are synonymous terms," for being
becomes loosed from its association with existence and comes to denote, in a phrase that will
be shortly introduced, the "habits of regularity" through which the world of objects finds its
meaningful expression in human cognition. But this is to anticipate aspects of Peirce's
thought​ ​which​ ​remain​ ​to​ ​be​ ​discussed.

Out of a sense of economy I addressed only four of the se ven questions posed in
Questions Concerning Certain Faculties Claimed For Man", however I believe the answers to
them effectively set the stage of Peirce's critique and show that the viability of the theory of
intuition in the explanation of cognition is at least called into question if not effectively
discredited. Peirce's own response to the implications in this paper are summed up in his
"Consequences of Four Incapacities", published in concert with the first and meant to further
propel​ ​the​ ​argument​ ​of​ ​his​ ​critique.

If the first paper can be seen as negative in character, the second is more positive.
Peirce takes it not only as an opportunity to draw out the implications of the first but to
provide a broad sketch of his conception of the nature of philosophical activity. He begins
with a brief look at how philosophical method shifted between the Scholastics and the
modern age initiated by Descartes. Peirce was widely read in the Scholastics and took much
acknowledged inspiration from them, especially Duns Scotus, for the formation of his
thought. By putting these two modes of thought side by side he hopes to throw into greater
relief the shortcomings of the Cartesian method and to thus clear the stage for the direction he
believes philosophy ought to pursue. His comparison is so concise as to defy paraphrase so I

23
quote​ ​it​ ​here​ ​in​ ​full.

.​ ​.​ ​.​ ​Cartesianism​ ​[.​ ​.​ ​.​ ​]​ ​may​ ​be​ ​compendiously​ ​stated​ ​as​ ​follows:
1. It teaches that philosophy must begin with universal doubt; whereas scholasticism
had​ ​never​ ​questioned​ ​fundamentals.
2. It teaches that the ultimate test of certainty is to be found in the individual
consciousness; whereas scholasticism had rested on the testimony of sages and of
the​ ​Catholic​ ​Church.
3. The multiform argumentation of the Middle Ages is replaced by a single thread of
inference​ ​depending​ ​often​ ​on​ ​inconspicuous​ ​premises.
4. Scholasticism had its mysteries of faith, but undertook to explain all created
things. But there are many facts which Cartesianism not only does not explain but
renders absolutely inexplicable, unless to say that "God makes them so" is to be
regarded​ ​as​ ​an​ ​explanation​ ​(5.264).

Peirce is not expressing a wish to return to scholasticism, but sees rather, in the contrast
afforded here, a manner in which modern science and logic can escape the difficulties of the
Cartesian problematic. Indeed, in the fourth statement above, the dilemma mentioned earlier
is clearly presented. In the Cartesian epistemology logical method is either ultimately
inexplicable and thus reduced to a description of those strategies of reasoning we find useful
to adopt, or is explained by saying simply that "God makes it so." Both are unpalatable
compromises for Peirce. Instead, what follows in "Some Consequences of Four Incapacities"
is his conception of a milieu of philosophical activity that will give logic the scope and
autonomy it requires if we are truly to understand the nature of knowledge and its growth in
the​ ​sciences.

Peirce takes each of the four points in turn. As to the characteristic skepticism of
Cartesianism, he regards it as self-deceiving. We cannot begin with complete doubt, he says.
The prejudices we carry with us in our reasoning cannot be discarded with the whim of a
maxim, because they do not occur to us as things that can be questioned. This observation
will be an important principle in his paper "The Fixation of Belief" in which he begins to
work out his logic of inquiry. "Let us not pretend to doubt in philosophy what we do not
doubt​ ​in​ ​our​ ​hearts."15

Peirce treats a similar theme in the second and third points - - the reasoning process and

15
​ ​Wiener,​ ​p.​ ​40.

24
the nature of the individual in it. Again he sees in the reification of the cogito, in this
conception of the individual self as a vehicle sufficient for the attainment of truth, the error
which has unnaturally focused philosophical attention on the problem of certainty and thus
warped the direction of philosophy from other more profitable courses of inquiry. Should we
all take this as a serious maxim of inquiry, metaphysicians would soon be dancing, with all
the angels in train, on a head of a pin, proclaiming metaphysics to have reached a point of
exactitude and certainty beyond the physical sciences. Although in agreement on this, they
would agree on nothing else. Invoking the method of the sciences Peirce says that we cannot
individually hope to attain truth, but only for the community of inquirers. Let "disciplined
and candid minds carefully examine a theory" so that the weight of it may ride or fall on the
acceptance​ ​of​ ​all.

It is clear that Peirce's scientific training comes to bear here. For him it is the methods of
the successful sciences which philosophy ought to imitate, to the extent that the terms of
philosophical study be removed from the nether world of the cogito and situated in the public
and observable course of inquiry engaged in by a community of inquirers. He champions the
"multitude and variety" of arguments in scientific inquiry as against the conclusiveness of
any​ ​one.​ ​"Reasoning,"​ ​he​ ​says,​ ​"should​ ​not​ ​form​ ​a​ ​chain​ ​which​ ​is​ ​no​ ​stronger

than its weakest link, but a cable whose fibers may be ever so slender, provided they are
sufficiently​ ​numerous​ ​and​ ​intimately​ ​connected."

In addressing the fourth point Peirce concludes by remarking once again on that
intransigent hallmark of nominalism, ushered in by Descartes and found in Kant and the
psychologists, that supposes some inexplicable, unanalyzable ultimate. Whether a benevolent
God or the thing-in-itself, Peirce desires to be rid of it. Philosophy should never admit of the
absolutely inexplicable, for as will be clear later, such a maxim would hinder the genuine
pursuit of inquiry. To put philosophy within this limit is as damaging and non-sensical as
starting off philosophy with an imagined doubt. Many of the problems of philosophy Peirce
sees, as Wittgenstein would later, as resulting from initial confusions over method and
perspective.​ ​This​ ​particular​ ​confusion​ ​is​ ​for​ ​Peirce​ ​one​ ​of​ ​the​ ​most​ ​pernicious.

His attempt to resolve the nominalist problem however begins in this paper and the
picture we see evolving here is an attractive one, of philosophy as a fallible, communal form
of inquiry. It may be argued that though Peirce makes some valid criticisms, he hardly

25
debunks Cartesianism as a method, nor much less offer proofs of his own theses. And the
argument would be right, except that Peirce is concerned with the very nature of inquiry and
method itself. The burden of "proof" lies not with him but with one who would hold that
philosophy is subject to the special constraints of Cartesianism. In fact, to seek in these early
papers on logic, inquiry, and belief anything approaching an axiom or an indubitable proof is
to remain engaged in a Cartesian framework and to miss the point of his critique. This is not
to say that he can simply ask us to accept an unwarranted series of assertions about the nature
of knowledge. As he says above, the "fibers" which will go to form the "cable" of his
philosophy, will, though slender, be intimately connected and will attempt to give
explanatory power to our experience of knowing. The terms of his philosophy are not given
axiomatically, but provisionally. All he can do is present a new picture and asks us to test it in
the course of inquiry, to see if what hitherto seemed to be problems resolve themselves in the
course​ ​of​ ​this​ ​new​ ​framework.

Having begun this new picture he augments it by returning to the first paper we
discussed​ ​and​ ​summarizing​ ​the​ ​conclusions​ ​to​ ​be​ ​drawn​ ​from​ ​it.

1. We have no power of Introspection, but all knowledge of the internal world is


derived​ ​by​ ​hypothetical​ ​reasoning​ ​from​ ​our​ ​knowledge​ ​of​ ​external​ ​facts.
2. We have no power of Intuition, but every cognition is determined logically be
previous​ ​cognitions.
3.​ ​We​ ​have​ ​no​ ​power​ ​of​ ​thinking​ ​without​ ​signs.
4.​ ​We​ ​have​ ​no​ ​conception​ ​of​ ​the​ ​absolutely​ ​incognizable​ ​ ​(5.265).

Whether these theses recommended themselves, one or all, in some other context, in the
course of his scientific work, or in some other philosophical investigation, it is clear that he
presents them here as the converse of the series of critiques in the first paper. They are the
consequences of capacities which Peirce hopes to have, if not disclaimed for man, at least
seriously called into question. With these claims Peirce provides the framework which will
give​ ​flesh​ ​to​ ​his​ ​study​ ​of​ ​the​ ​course​ ​of​ ​inquiry​ ​and​ ​the​ ​subsequent​ ​formation​ ​of​ ​his​ ​semiotic.

Though "Consequences of Four Incapacities" continues at length, the discussion is more


or less a repeated analysis of the critique in the first paper. However, what can be profitably
taken in concluding a look at these two papers is the new conception of mind Peirce hints at.
He wants to break open the occult regions of the cogito in which mind is thought to inhere

26
and to show that it is not something intuitive and private, but rather public, and interactive.
His conception of mind will be a phenomenon demonstrable in the practices of inquiry, a
reality not cut off from the world of its knowing but a constitutive part of it. The picture of
philosophy and its method he gives at the beginning of this second paper is instructive as well
for what will follow, especially the notion of its being a collaborative enterprise in the efforts
of​ ​an​ ​engaged​ ​community​ ​of​ ​inquirers.

Running through all of this is the unmistakable influence of his experience as a scientist,
the sense in which, through a particular method of investigation, the objects of science
achieve a clarity hitherto unfound in philosophical research. Philosophy will gain its proper
footing, he says, if only it can approximate the method of the successful sciences. It is not an
unwarranted statement. The question of method will be of primary importance in his
examination​ ​of​ ​inquiry​ ​in​ ​"The​ ​Fixation​ ​of​ ​Belief",​ ​to​ ​which​ ​I​ ​know​ ​turn.

The​ ​Fixation​ ​of​ ​Belief

To recall, his examination of inquiry, and the subsequent formulation of cognition as a


semiotic process, have at their base a concern for the elucidation of a new logic. It was his
ultimate dissatisfaction with Kant's logic that served, as I have stated, as an impetus for his
early philosophical study. In dismantling Kant's epistemology, he hopes to give in the one
that he substitutes a logic adequate to the building of a new architectonic, one that in its
autonomy provides the scope that cognition needs if our understanding of its processes are
truly​ ​to​ ​aid​ ​us​ ​in​ ​the​ ​pursuit​ ​of​ ​truth.

The logic he substitutes takes as its model the role of inquiry in the natural, empirical
sciences. If philosophy should approximate the sciences, it should be of concern the manner
in which scientific inquiry achieves its results. In examining this process Peirce hopes to
show​ ​a​ ​connection​ ​between​ ​this​ ​empirical​ ​logic​ ​and​ ​the​ ​way​ ​the​ ​mind​ ​works.

In the course of scientific inquiry reasoning moves from what is known to what is not
yet known. In traversing this gap Peirce discerns four assumptions which condition its
success.

l.​ ​Inquiry​ ​presupposes​ ​a​ ​guiding​ ​principle.

27
2.​ ​The​ ​settlement​ ​of​ ​belief​ ​is​ ​best​ ​achieved​ ​by​ ​good​ ​reasoning.
3.​ ​The​ ​use​ ​of​ ​the​ ​scientific​ ​method,​ ​which​ ​is​ ​the​ ​proper​ ​method​ ​of​ ​good​ ​reasoning.
4. The hypothesis that there must be a reality that serves as the justification for
pursuing​ ​inquiry.

Essentially these assumptions can be boiled down to the second in the list, that the aims of
inquiry are best achieved by good reasoning. The other three actually describe the framework
and​ ​the​ ​means​ ​by​ ​which​ ​reasoning​ ​functions​ ​well.

Basic to all reasoning is the presumption of a guiding principle, or what Peirce


describes as a habit of mind. It is a proposition which formulates a general rule of inference.
When we infer something and thus give a conclusion to a premise, it is not done willy nilly,
but rather under the force of a guiding constraint. "All men are mortal" is a principle we use
to constrain the inference that if "Socrates is a man" then "Socrates is mortal." The exercise
of such a principle is vital to reasoning, for if we began inquiry without the hope of the
resolution of doubt that such principles offer, we would live in a continuing state of doubt
with​ ​no​ ​predisposition​ ​to​ ​engage​ ​the​ ​world​ ​in​ ​any​ ​specific​ ​way.​ ​Our​ ​lives​ ​would​ ​be​ ​chaotic.

We employ many such principles in the course of our reasoning, but for Peirce some are
more fundamental than others and play an essential role in all reasoning, for instance those
which guide our questions about the very process of reasoning itself. It is his concern here to
determine just what those are, for if we use true guiding principles then our reasoning will be
sound. The question is how this is to be done, how to argue for the truth of one principle over
another. He says, " . . .those rules of reasoning which are deduced from the very idea of the
process are the ones that are most essential" (5.369). The answer lies in considering the
assumptions inherent in the very questions that logic asks, presuppositions without which the
questions of logic would not even arise. One such assumption or guiding principle, with
which Peirce will form the backbone of his conception of inquiry, is revealed when we
consider "why a certain conclusion is thought to follow from certain premises" (5.369). The
principle assumed in asking this and which constrains us to conclude it is "that there are such
states of mind as doubt and belief -- that a passage from one to the other is possible, the
object of thought remaining the same, and that this transition is subject to some rules by
which all minds are alike bound." This is one among those few principles which have a
fundamentally essential relevance in the broad course of inquiry and is what Peirce calls the

28
doubt belief principle. This principle is singled out in his analysis as being a necessary
element in the process of inquiry, at least if we are to be conscious of inquiry reaching its
aim.

Thus, if inquiry is to be successful, and good reasoning is the means by which that is
accomplished, we must take as an essential characteristic of reasoning that it's inception and
terminus be defined in the passage we take from a state of doubt to one of belief. I will
discuss shortly what Peirce means by 'doubt' and "belief but I want first to re-emphasize the
over-all motivation of his investigation here. It certainly concerns the elucidation of the
structure of inquiry and how it is that it achieves its results, but more importantly it is to show
what principles of thought an inquirer should adopt in any particular inquiry. What are the
aims​ ​that​ ​guide​ ​our​ ​reasoning?

The answer is somewhat complex, and is reflective of the tension I spoke of at the
beginning of this paper. On the one hand Peirce is concerned with how inquiry should be
conducted, with determining those principles we should adopt if our inferences are to arrive
at true conclusions. On the other hand, however, he is concerned that the practice of inquiry
not be ignored as something secondary and of minor concern. It is his vision of philosophical
inquiry in a communal context that will give the characteristic emotional and physical stages
in​ ​inquiry​ ​a​ ​major​ ​role​ ​in​ ​the​ ​formation​ ​of​ ​meaning​ ​itself.

So, with regard to the practical process of inquiry itself, Peirce defines its aim as
nothing other than the settlement of opinion. Whether the truth be gained in reality or not is
immaterial for understanding how inquiry works as such. As long as some belief has been
reached inquiry has fulfilled its aim. This returns us to Peirce's doubt-belief principle. Inquiry
is initiated when we begin to experience doubt about something, when whatever particular
system of beliefs we have fails our expectations about the world. A belief is similar to a
guiding principle in that it can be seen as a regularity in thinking. Behaviorally it manifests
itself as a disposition to act in accordance with the consequences we expect from holding that
belief in the world. For example the belief that "standing in the rain will make me wet"
disposes me to stay inside (if it is in fact my desire to stay dry). If I were to find that I
remained​ ​dry​ ​while​ ​standing​ ​in​ ​the​ ​rain,​ ​I​ ​would​ ​begin​ ​to​ ​experience​ ​doubt​ ​about​ ​that​ ​belief.

In "The Fixation of Belief" the word 'belief is very quickly subsumed by the word

29
"habit and in the course of all his subsequent writing comes to represent for Peirce a central
characteristic of the inquiring knowing subject. A full exposition of it will come later in my
discussion of his pragmatism and semiotic but for now it is sufficient to hint at the
importance it will come to hold in Peirce's thought. The words 'belief or "habit' are terms
which have as their object something Peirce calls a general. A general is something that
persists between different acts of conceptual awareness, something repeatable, and can be
compared to what has elsewhere in the tradition been understood by the Aristotelian term
'universal'. A general is not, in the Platonic sense, a static individual thing but is rather
dynamic and evolves as the process of inquiry moves forward. As Hausman has described,
" . . . generals are what are referred to by expressions that indicate repeatability of the
referent."16 Thus, a habit is not an individual instance of behavior but rather is a rule which,
by virtue of its repeatability, relates the aspects of the world of our experience in a
meaningful way. To put it simply, habits of thought are the threads by which the tapestry of
our knowledge of the world is woven. They are not merely convenient devices for ordering
the world, as a nominalist would claim, but are actual representations (though approximate)
of​ ​the​ ​lawful​ ​patterned​ ​relation​ ​cohering​ ​in​ ​the​ ​things​ ​of​ ​the​ ​world.

To return to the doubt-belief principle, Peirce describes the feeling of doubt as an


irritation, an uncertainty that manifests itself in blocking our willingness to act. In this sense
it is something decidedly different from the unreal hypothetical doubt of which he accused
Descartes. It was the point of his critique there to say that an understanding of the reasoning
process cannot be severed from the very real conditions which give rise to it. So with the
disruption of belief we begin to feel the irritation of doubt and seek to resolve it with the
establishment of a new belief. In this basic but remarkable formulation Peirce takes the
cognitive process, hitherto occluded in the Cartesian cogito, and exposes it in a manner
amenable to the kind of readily observable analysis by which scientific investigation
proceeds. In fact it will be our conscious awareness of the characteristic stages of inquiry
which will give Peirce's pragmatic maxim real operative power in the determination of
meaning.

16
​ ​Carl​ ​R.​ ​Hausman,​ ​Charles​ ​S.​ ​Peirce's​ ​Evolutionary​ ​Philosophy.​ ​(Cambridge:​ ​Cambridge​ ​University​ ​Press,​ ​1993)​ ​p.​ ​26.

30
This does not conclude Peirce's analysis of inquiry however. The disruption of a habit
of thought can be an indication of more than just the frustration of a particular goal. More
deeply it may be a sign that our conception of the world is not right. And so the tension about
which I spoke earlier returns. Why should an inquirer be concerned about the proper
conception of the world if his or aim is simply to settle opinion, to establish a new belief
which​ ​will​ ​appease​ ​the​ ​irritation​ ​of​ ​doubt​ ​and​ ​allow​ ​him​ ​or​ ​her​ ​to​ ​carry​ ​on?

The answer lies in a further presupposition of inquiry and a consideration of the proper
method​ ​to​ ​adopt​ ​in​ ​inquiry​ ​in​ ​light​ ​of​ ​the​ ​claim​ ​made​ ​by​ ​this​ ​presupposition.

For Peirce one of the presuppositions of logic is the "Hypothesis of Reality." As he


says,

There are Real things, whose characters are entirely independent of our opinions
about them; those realities affect our sense according to regular laws, and, though
our sensations are as different as our relations to the objects, yet, by taking
advantage of the laws of perception, we can ascertain by reasoning how things really
are; and any man, if he have sufficient experience and reason enough about it, will
be​ ​led​ ​to​ ​the​ ​one​ ​true​ ​conclusion​ ​(5.384).

As formulated here Peirce presents a straightforward position of realism, one in which the
Real is not dependent upon any individual opinion but rather, through its action on
perception, affects opinion, an opinion which will be the object of a consensus among people
who​ ​have​ ​enough​ ​experience​ ​and​ ​conduct​ ​their​ ​inquiries​ ​correctly.

As an assertion of actual fact its warrant may be questioned. but as Hookway observes,
"The question whether the hypothesis is true is distinct from that of whether it is a
presupposition of logic."17 It is the latter that Peirce believes to be true. To recall, "The
Fixation of Belief" is concerned with finding the aims that guide inquiry. Though the
doubt-belief principle is one such guide, it serves more in a structural, functional way, much
like railroad tracks that guide a train along its course. What Peirce is looking for, to extend
the metaphor, is something that guides inquiry to a final destination, something that chooses
between two tracks that diverge at a fork. It is clear that we makes such decisions. The
question is, what presupposition of logic constrains us in making them? To say that "the
truth" is this guide is too narrow, for Peirce observes that when opinion is settled, whatever it

17
​ ​Christopher​ ​Hookway,​ ​Peirce​.​ ​(London:​ ​Routledge​ ​&​ ​Kegan​ ​Paul​ ​pic,​ ​1985)​ ​p.44.

31
is we believe, we believe per force to be true. What he needs is something not affected by
opinion,​ ​and​ ​it​ ​is​ ​the​ ​"Hypothesis​ ​of​ ​Reality"​ ​that​ ​does​ ​just​ ​that.

But is this hypothesis a presupposition of logic? Peirce answers this by examining in


"The Fixation of Belief" four methods by which inquiry can proceed in settling questions. In
brief they are: (1) the method of tenacity, (2) the method of authority, (3) the a priori method,
and (4) the scientific method. If inquiry is begun to appease the irritation of doubt, Peirce will
show that the adoption of only one of these methods will make that aim possible and that this
method​ ​will​ ​have​ ​incorporated​ ​into​ ​it​ ​the​ ​hypothesis​ ​of​ ​reality.

Peirce begins with the method of tenacity. As its name suggests this method is
characterized by the dogged holding of a belief sustained in disregard of any other mediating
factors. Superficially it answers, most immediately and pointedly, the directive implied in
inquiry's​ ​aim​ ​-​ ​the​ ​settlement​ ​of​ ​belief.

This method is unacceptable however. If we are simply to acquire a new belief, why
not, Peirce asks, "[take] as answer to a question any we may fancy, constantly reiterating it to
ourselves, dwelling on all which may conduce to belief, and learning to turn with contempt
and hatred from anything that might disturb it?" (5.377). His answer is that the social impulse
is against it. Unless we lives as hermits, our lives and the beliefs which guide them will
intersect with those of others. We will see others holding different opinions and be forced to
reconsider the confidence we hold in our own. The aim of inquiry deepens in our experience
of the community in that the belief which removes doubt must endure not only the experience
of the individual but that of the community as well. In order to endure as long as possible,
"the problem becomes how to fix belief, not in the individual merely, but in the community"
(5.378).

The next logical step then in the settlement of questions is what Peirce calls the method
of authority. As Hobbes answered the problem of a life nasty, brutish, and short with the
reification of the state, so here the settlement of opinion is transferred to a political or social
body​ ​with​ ​authority​ ​over​ ​all.

Though the fixation of belief is more enduring by this method, the problem of the first
persists. Settlement is still achieved by a tenacious whim that cannot hope to secure itself
against variations of opinion on some subject, whether within the realm of its authority or
outside it, in a different land, or a different historical time. Soviet Russia in this century

32
would be the most characteristic example of this. Again, Peirce says that the social impulse is
against it. In Soviet Russia it was the intellectual class - the writers, scientists, and
philosophers who, in the persistency of their questions, made clear the ultimate inadequacy of
the​ ​method​ ​of​ ​authority.

From this class, however, a third method suggests itself -- the a priori. It is an
appropriation of characteristics of the first two in that the authority returns from the state to
the individual, however it is no longer an idiosyncratic will which directs inquiry, but the
authority of reason. By centering the authority to settle opinion in a capacity equal to and
accessible by all, the effective scope of belief moves beyond the boundaries of the state and
encompasses the entire community of inquirers. This method seems to have much to
recommend it. It avoids the idiosyncratic character of the first two while bringing the
permanence of belief into the realm of real possibility by employing "the light of reason", a
capacity common to all. The a priori method comes thus far closest to satisfying the aims of
inquiry.

Peirce would almost agree, given the more catholic scope of this method in inquiry, but
he sees again as argument against it the power of the social impulse. As the history of
Western thought readily shows, that which seems "agreeable to reason" has changed in
different times and places and is thus more a function of taste than anything else. It is "always
more or less a matter of fashion, and accordingly metaphysicians have never come to any
fixed agreement, but the pendulum has swung backward and forward . . ." (5.383). It is
inevitable that a belief which claims to follow the light of reason will be found inadequate to
the experience of some. Their resistance to it will throw not only individual beliefs into
question, but the very method which supports them. This continuing pressure of the social
impulse is what leads Peirce to remark, "To satisfy our doubts, therefore, it is necessary that a
method should be found by which our beliefs may be determined by nothing human, but by
some​ ​external​ ​permanency​ ​-​ ​by​ ​something​ ​upon​ ​which​ ​our​ ​thinking​ ​has​ ​no​ ​effect"​ ​(5.384).

The other three suffer from dependency on some aspect of human will and the
limitations inherent in human finitude. Some agency ultimately external to human demands
would provide the conditions necessary for the more enduring, if still theoretically
provisional, establishment of belief. It is a method which "must be such that the ultimate
conclusion of every man shall be the same. Such is the method of science" (5.384). This

33
method is the only viable one for Peirce because fundamental to its activity is its hypothesis
of reality, that there are "Real" things that affect our senses, and thus formulate our beliefs,
according to regular laws. The hypothesis serves as a touchstone to our idiosyncrasy and
provides the only real possibility of the settlement of opinion, even if only in principle or in
the​ ​long​ ​run.​ ​And​ ​the​ ​only​ ​method​ ​that​ ​engages​ ​this​ ​hypothesis​ ​is​ ​the​ ​scientific​ ​method.

So if one of the conditions of achieving the aims of inquiry is good reasoning, then the
method of science provides the means by which that is accomplished. However this method
is only plausible if the presupposition, which serves as the criterion for its choice, is justified.
Must​ ​the​ ​realist​ ​hypothesis​ ​be​ ​assumed​ ​if​ ​inquiry​ ​is​ ​to​ ​achieve​ ​its​ ​ains?

This fourth of the conditions of inquiry enumerated above, that the hypothesis of reality
is necessary in order for inquiry to proceed, is a fundamental feature of Peirce's philosophy,
one elaborated over the course of the years in successive stages of his thought. Its treatment
here in this early paper though is brief, and as Peirce admitted, not yet entirely defended. As
he says, "If this hypothesis is the sole support of my method of inquiry, my method of inquiry
must not be used to support my hypothesis" (5.384). He does offer some points in its support
however. First, that this method is widely used, everyone using it most of the time, which
shows that no one seriously doubts it. And second, that its use has only demonstrated its
effectiveness in aiding inquiry, rather than engendering doubts about it. He continues by
noticing how the practice of the three other methods sows doubts in the minds of their users
whereas​ ​use​ ​of​ ​the​ ​scientific​ ​method​ ​only​ ​bolsters​ ​the​ ​confidence​ ​of​ ​its​ ​practitioners.

These points are admittedly weak because they rest on inductive experience and cannot
really lay claim to questions of logic. A fourth point however suggests more deeply the point
of Peirce's argument in this paper. It returns to the initiation of inquiry in the irritation of
doubt. We are prompted to make inquiry when we find our beliefs to be inconsistent, when
they clash with our experience. Implied in this is that we find untenable a situation where
differing propositions about the same thing are true. " . . . but here already is a vague
concession that there is some one thing to which a proposition should conform." The realist
hypothesis must be assumed in inquiry, and thus be a presupposition of logic, for otherwise
there​ ​is​ ​no​ ​means​ ​to​ ​explain​ ​the​ ​motivation​ ​to​ ​remove​ ​the​ ​inconsistencies​ ​in​ ​our​ ​experience.

Again Peirce is not asking whether the hypothesis is true, but only if it is a logical

34
assumption of inquiry. As Hookway observes, "starting from a characterization of the aim of
inquiry as the settlement or fixation of belief, Peirce hopes to derive a statement of the
'presuppositions of the logical question' from a consideration of how certain methods of
inquiry simply cannot be adopted. Simply asking "How should I conduct my inquiries?"
reveals my acceptance of propositions which help me to answer the question".18 It is in this
sense that the inadequacy of the three other methods reveal themselves. In their failure to
accept the hypothesis of reality, the logical question does not arise. To use them is to remain
oblivious to the question of method at all, except insofar as the doubt they engender compels
one to search for a better method, one like that of science, which justifies itself in the
assumptions​ ​of​ ​the​ ​very​ ​questions​ ​it​ ​asks.

The upshot of the "Fixation of Belief" is that the scientific method is the only one which
can be self-consciously used in aiding the aims of inquiry. Its assumption of the hypothesis of
reality makes real the possibility of settled belief because, though belief in the short run
remains provisional, it provides the confidence that in the course of inquiry our doubts are
assuaged​ ​not​ ​by​ ​whim​ ​or​ ​idiosyncrasy​ ​but​ ​by​ ​some​ ​external​ ​permanency.

This picture of inquiry remains incomplete however. It began with a very


instrumentalist description of its process, where the aim was nothing other than the settlement
of opinion regardless of considerations of truth, and ended by pointing inquiry toward the
normative aims contained in the hypothesis of reality. Though Peirce adduced a number of
strongly common sense arguments in justifying this move, a closer look at the structural
elements of inquiry is necessary in order to understand how his conception of inquiry
provides a logic which gives a normative sense to the broader aims implied in the hypothesis
of​ ​reality.

Thus far inquiry has been described in terms of the structural roles played by doubt and
belief and has been located in the context of a community of inquirers where in the result of
whose deliberation a final opinion is hoped to be found in the long run. But what
characterizes this deliberation more concretely? In addition to the possibility of attaining one
final opinion, Peirce sees the indispensability of the scientific method in its conformity to the
laws of inference. Inference, it will be recalled, is the basic unit of cognition Peirce replaced
with intuition. It is the means by which faulty beliefs are replaced and by which inquiry

18
​ ​(Hookway,​ ​p.​ ​49)

35
reaches​ ​its​ ​aim.​ ​It​ ​is​ ​the​ ​inferential​ ​nature​ ​of​ ​thinking​ ​which​ ​needs​ ​a​ ​closer​ ​examination.

Traditionally, logicians have viewed the various forms of syllogistic argument as being
structural expressions of the actual activity of thinking. As Gallie has demonstrated,19 in the
act of inferring, someone starting with a known truth finds herself induced to conclude some
other truth not known beforehand. The traditional understanding of what happened in such a
case was that in the act of thinking there occurred a unique passage or movement of thought
characterized in the following way: a) a given truth is assented to and held as a premise, b)
something about the particular arrangement of the premise or premises recommends in a
forceful way a further truth. In stating this new truth in the form of a conclusion, the
inference​ ​comes​ ​to​ ​an​ ​end.

Thus was the act of thinking conceived, and it was the purpose of argument, whether
written or oral, to express this -- and not only this, but in addition to create in the mind of a
hearer or reader the same movement of thought which led to the conclusion. Though there be
some trace of common sense in this view, Peirce rejects it almost completely. I make these
remarks because it provides an opportunity to re-emphasize the radically different conception
of mind Peirce has, not only as it is given in his understanding of inference, but in inquiry as
a​ ​whole,​ ​as​ ​will​ ​be​ ​seen,​ ​in​ ​his​ ​semiotic.​ ​Regarding​ ​inference​ ​Peirce​ ​says,

Practically, when a man endeavors to state what the process of his thought has been
after the process has come to an end, he first asks himself to what conclusion he has
come. That result he formulates in an assertion, which, we will assume, has some
sort of likeness . . . with the attitude of his thought at the cessation of the motion. . . .
he next asks himself how he is justified in being so confident of it; and he proceeds
to cast about for a sentence expressed in words which shall strike him as resembling
some previous attitude of his thought, and which at the same time shall be logically
related to the sentence representing his conclusion, in such a way that if the
premise-proposition be true, the conclusion-proposition necessarily or naturally
would​ ​be​ ​true​ ​(2.27).

Essentially Peirce is saying that in talking about an inference we in no way refer to some
process having transpired in the mind, but simply that earlier beliefs are available to confirm
new ones. An inference represents not a particular movement of thought but rather that given
a proposition or belief that we newly hold, we are in a position to give a reason for it. For
Peirce there may be a hundred different ways of thinking in passing from a premise to a

19
​ ​W.B.​ ​Gallie,​​ ​Peirce​ ​and​ ​Pragmatism.​ ​(New​ ​York:​ ​Dover​ ​Publications,​ ​Inc.,​ ​1966)​ ​p.95.

36
conclusion. "Inference . . . may be, not to say probably is, of an entirely different construction
from​ ​the​ ​thinking​ ​process"​ ​(2.59,​ ​cf.​ ​2.183).

All this, again, to re-emphasize Peirce's departure from the Cartesian tradition.20
Whatever the mental process may be, understanding it as such is of no practical importance in
making an inference and using the results of inquiry. And though inquiry itself was initially
described in terms of doubt and belief -- terms we associate with mental states -- neither
should it be restricted to just a kind of thinking. For Peirce it covers any kind of activity,
physical or mental, which removes the irritation of doubt. And though it has yet to be
discussed, Peirce's analysis of sign activity will be given a similar characterization and will
enlarge to a wide degree his insistence that "it is not to private processes of thought, but to
arguments, statements, and terms which figure in public discourse, that logical principles are .
.​ ​.​ ​applied"​ ​(Gallie,​ ​p.92).

But to return to inference, though Peirce classes under the rubric of the laws of
inference the traditional modes of deduction and induction, he extends its scope to include
what is commonly known as hypothesis, or as Peirce terms it, abduction. Of deduction and
induction Peirce offers traditional analyses, but as regards induction he makes some original
and important observations. Rather than originating new knowledge by advancing from what
is already known, he sees induction as being primarily a method of testing. In deduction we
discern those conclusions which must hold if the premise (a particular hypothesis) is true. To
use​ ​an​ ​example​ ​of

Peirce's:

Most​ ​people​ ​wounded​ ​in​ ​the​ ​liver​ ​die.


A​ ​was​ ​wounded​ ​in​ ​the​ ​liver.
So,​ ​probably​ ​A​ ​will​ ​die.

Induction tests this conclusion against experience and in the long run either verifies it or finds

20
​Peirce's characterization of Descartes is confirmed by a nuance found in Descartes where he says, "So I shall run through
them several times in a continuous movement of the imagination, simultaneously intuiting one relation and passing on to the
next, until I learn to pass from the first to the last so swiftly that memory is left with practically no role to play, and I seem to
intuit the whole thing at once." Descartes. Rules for the Direction of the Mind, in The Philosophical Writings of Descartes​,
trans.​ ​John​ ​Cottingham,​ ​et,​ ​al.,​ ​(Cambridge:​ ​Cambridge​ ​University​ ​Press,​ ​1985),​ ​Vol.​ ​I,​ ​p.​ ​25.​ ​AT​ ​388).

37
an instance which disproves it and thus it is discarded. It is this capacity of inference which
for Peirce recommends only more forcefully the scientific method, that is, its self-correcting
nature. By testing belief against experience induction furnishes a method by which
idiosyncrasy​ ​is​ ​minimized​ ​and​ ​consensus​ ​of​ ​opinion​ ​increased.

But as was stated, induction does nothing to advance knowledge, but serves only to
further sediment belief. What introduces a new "premise" into the arena of inquiry is a form
of inference Peirce called abduction. Its activity is clearly distinct from induction as he shows
in an example, " . . . any historical fact, as that Napoleon Bonaparte once lived, is a
hypothesis; we believe the fact, because its effects -- I mean current tradition, the histories,
the monuments, etc. - are observed. But no mere generalization of observed facts induction)
could teach us that Napoleon lived" (2.714). The distinction between the two are even more
decisively​ ​drawn​ ​in​ ​the​ ​following​ ​passage,

By induction, we conclude that facts, similar to observed facts, are true in cases not
examined. By hypothesis, we conclude the existence of a fact quite different from
anything observed, from which, according to known laws, something observed
would necessarily result. The former, is reasoning from particulars to the general
law;​ ​the​ ​latter,​ ​from​ ​effect​ ​to​ ​cause.​ ​The​ ​former​ ​classifies,​ ​the​ ​latter​ ​explains​ ​(2.636).

Both induction and abduction are ampliative forms of reasoning but the latter for Peirce is the
real catalyst of progress in inquiry. To return to his rewording of Kant, discovering how
synthetic reasoning is possible at all is the lock upon the door of philosophy, and here Peirce
begins​ ​to​ ​turn​ ​the​ ​key.

The abductive inference forms a hypothesis to account for some surprising fact in
experience which eludes explication by facts already known. As Gallie illustrates in the
following​ ​schema:

1)​ ​The​ ​surprising​ ​fact,​ ​C,​ ​is​ ​observed.


2)​ ​But​ ​if​ ​A​ ​were​ ​true,​ ​C​ ​would​ ​be​ ​a​ ​matter​ ​of​ ​course,
3)​ ​Hence,​ ​there​ ​is​ ​reason​ ​to​ ​suspect​ ​that​ ​A​ ​is​ ​true.

Whether or not it is a viable hypothesis depends on further inferences, deducing the


consequences of the hypothesis and testing them inductively in the course of experience.
Abduction alone cannot account for the growth of knowledge but only in the mutual

38
dependence it shares with the other two forms of inference. It remains to be understood
however how abduction furnishes hypotheses pertinent and valuable to the inquiry at hand.
Any hypothesis will open up roads of inquiry, but only one or a very few will explain the
facts in any given situation. How, from the seemingly infinite number of hypotheses, are we
to​ ​choose​ ​one​ ​worthy​ ​of​ ​inductive​ ​testing?​ ​Peirce​ ​answers​ ​in​ ​the​ ​following​ ​way,

Science presupposes that we have a capacity for 'guessing' right. We shall do better
to abandon the whole attempt to learn the truth . . . unless we can trust to the human
mind's having such a power of guessing right that before very many hypotheses
shall have been tried, intelligent guessing may be expected to lead us to the one
which will support all tests, leaving the vast majority of possible hypotheses
unexamined​ ​(6.530).

At first this response may sound like special pleading, being more descriptive than
explanatory in nature. Descriptively it is correct. We don't flail about in inquiry grasping
indiscriminately at whatever hypothesis may come our way. As the history of science shows
we come very quickly to narrow the choice of hypotheses down to those few which have
explanatory​ ​pertinence.​ ​But​ ​what​ ​explains​ ​this​ ​success?

For Peirce "the human mind is akin to truth" (7.220), which though it may seem vague
to say, is nothing more really than a startlingly concise statement of his epistemological
vision. As will be elaborated most fully in his semiotic, mind, and the reality which it
investigates, have fundamental structural similarities. The habits which guide thought are of
the nature of those habits, those regularities or what we call laws, through which the real is
manifested. As the history of science shows, in the early history of thinking, especially before
the widespread use of the scientific method, hypotheses which led to real discoveries came
few and far between, but as the constellation of more or less perdurable beliefs has grown and
a certain core of habits of inference become more stable and refined, insights have come
more​ ​readily.

In light of this illustration of the forms of inference Peirce's characterization of belief as


habit is perhaps more clearly understood. The habits Peirce speaks of are really habits of
thought, and as we have just seen the process of thought is the interplay of the forms of
inference, so that what Peirce means in speaking of a belief is really a habit of inference. The
life of thought is the formation and exercise of habits of inference. In fact it is for this
fundamental operation of the cognitive mind that Peirce formulates the pragmatic maxim.

39
Though I anticipate a later discussion of his pragmatism it may be said here that he conceived
of it as serving as a logical aid in the admissibility and formation of hypotheses. This point, to
be elaborated in the following chapter, provides further refinement to what Peirce perceives
as​ ​our​ ​already​ ​potent​ ​abductive​ ​abilities.

The​ ​Real​ ​and​ ​the​ ​Critical​ ​Control​ ​of​ ​Inference

I want to close this chapter with some further reflection on the notion of habit that will
draw out a few final implications of his theory of inference and lead to a discussion of his
semiotic. In general we mean by the word "habit something done unconsciously, without
thinking, something done by rote. That the experience of drawing inferences is not analogous
to the schema of a syllogism is implied in that common sense understanding , of habit and is
what Peirce meant in his criticism of traditional conceptions of thinking, We don't plod along
in a methodical syllogizing but rather we find ourselves "jumping" to conclusions. It is only
afterward, in defending a position, that we may find occasion to argue consciously with
reasons. This fluid, seemingly effortless jumping to conclusions is effected by the fact that
our beliefs, our guiding principles, reside in a habitual state. Their effectiveness in explaining
experience has woven them into the guiding patterns of thought which move inquiry forward,
and thus we find ourselves aware in the experience of thinking not so much of individual
strands,​ ​but​ ​of​ ​the​ ​pattern​ ​as​ ​a​ ​whole.

With regard then to the "unthinking" character of habits of inference, and to understand
the real force behind what he means by this term, I turn to a further threefold distinction
among inferences that Peirce makes in his theory. First, he speaks of what he calls
"Reasonings." These are the consciously articulated inferences of someone like a scientist or
a mathematician. The series of premises leading to a conclusion is elaborated with specific
intention and with a clarity sufficient to the aims of the investigation. Though he was critical
of this model insofar as it exemplified the traditional subject-matter of logic, he does admit of
the important role it plays in the life of the mind, for it is to a large extent in such explicit acts
of inference that beliefs come under the critical scrutiny capable of evaluating and reforming
them.

40
However, the great majority of our inferences are not of this fully articulate character.
When we have a birthday for example we realize that we are one year older than we were on
our last birthday, and we don't have to consciously syllogize the argument to see that it is so.
Or realizing that we will all one day die. The leading premises in the inference have become
so habitual through the course of experience that they have become in effect unconscious.
Even the mathematician formulating premises in the course of a proof; though she be
specifically conscious of the elements of his argument, and of the inferences which lead from
one premise to the next and finally to the conclusion, even in this process there is a great
many unconsciously habitual inferences made which support the explicit ones she is making.
These kinds of inferences Peirce calls a-critical, by which he means that they are not subject
to logical control. It is certainly possible to subject them to a Cartesian scrutiny, but the point
is that it does not occur to us to do so. In principle however they are corrigible, for if in the
course of experience they should lead us into error we can reform it in the formation of a new
hypothesis.

Finally, in addition to articulate and a-critical inferences, Peirce claims that there are
"operations of the mind which are logically exactly analogous to inference excepting only
that they are unconscious and therefore uncontrollable and therefore not subject to logical
criticism" (5.440). By saying 'operations of the mind' what he has in mind is our perceptual
judgments, for example that we see the sky as blue or that we see three apples on the table.
Though a - critical inferences are not as a matter of course given over to logical scrutiny in
the course of inquiry, they can in fact be made so subject, but perceptual judgments cannot. If
we see that an apple is red, we are psychologically incapable of conceiving it otherwise, of
thinking, while we are perceiving it, that we are mistaken in the judgment. Later we may
discover a set of particular conditions which led us to an erroneous judgment but, as such, we
cannot undo perceptual judgments. There is a givenness about then which is in essence
incorrigible.

The interesting point though is that Peirce sees perception as a form of inference, as
being "logically exactly analogous' to it. It is to abduction or hypothesis that he is referring
here. In his criticism of Descartes' Intuitionism we recall that what seemed most like the
result of intuition, namely perception, was really the outcome of a series of inferences, as he
demonstrated with examples of sight, the perception of texture, the hearing of a tone, etc.

41
Though perception seems like a brute unmediated given it is really analogous to hypothetic
inference. In addition to the examples above Peirce finds confirmation of this in the fact that
every perception, like every hypothesis, involves necessary consequences. With a hypothesis
we deduce its consequences and then attempt to confirm them inductively in experience. To
show that perception is analogous he uses an example. If one perceives that an event C is
subsequent to an event A, one must maintain as a consequence that any other event standing
in subsequence to C must have the same relation of subsequence to A as well. That this is so
seems patently obvious but Peirce says that we're justified in making this inference only if the
perception of the relationship between C and A "embodies all the necessary or defining
properties of the relation of apparent subsequence" (5.157). The point here is that we don't
apprehend these conditions in an articulate way, and so the question is how is it to be
explained that perceptual judgments have necessary consequences if there is no clear
mechanism​ ​for​ ​making​ ​them​ ​manifest?

Peirce answers that we can infer these consequences because the premises which would
allow us to draw them reside in our thinking habitually. It is this truly embedded and
unconscious working of inference in perception that makes it seem unmediated and forced
upon​ ​us​ ​and​ ​which​ ​accounts​ ​for​ ​our​ ​being​ ​unable​ ​to​ ​"undo"​ ​perception,​ ​to​ ​criticize​ ​it.

But this characterization of perception raises a problem. If thinking proceeds in the


inferential way he has described, by the formation and exercise of habits of inference, and if
our method of inquiry supposes a Real which constrains these habits, we must be able to
explain how the Real so constrains them. How is the reformation of habit actually
occasioned? That we constantly experience this constraint in daily life is obvious, yet Peirce's
account of cognition, and more pointedly of perception, as inferential, seems to deny the
possibility of a wholly new and first premise entering and constraining thought. If perception,
as the interface between thought and the Real, is analogous to hypothetical inference, how
does​ ​the​ ​"material"​ ​of​ ​hypothesis​ ​enter​ ​the​ ​process?

Peirce's answer lies in his conception of unconscious inferences. On the one hand the
unconscious habitual inferences which underlie perception make it theoretically liable to
error. If the premises are wrong so the conclusion will be as well, and so it might be
concluded that such an understanding of perception would be inadequate to Peirce's
constraining conception of the Real. At the same time however we have seen that perception

42
is not subject to logical scrutiny. We can control our reasoning and even much of the
a-critical inferences which guide the bulk of our mental life, but we cannot control how we
perceive. Our perceptual judgments are in a real way forced upon us and so, in the sense then
that we cannot criticize them because unconscious, they act as first premises. For all intents
and​ ​purposes​ ​perception​ ​does​ ​provide​ ​the​ ​grist​ ​for​ ​the​ ​inferential​ ​mill​ ​of​ ​our​ ​thinking.

The real problem above is not so much if perception is the real point of our contact with
the 'outward clash' of reality but if, as inferential, it can provide the occasions by which habits
of inference can be reformed. I think his answer shows that it can, however it also raises a
final difficulty which concerns a theme fundamental to his entire recasting of epistemology in
inferential terms. The question is, how can Peirce's position thus far stated elude a
fundamental skepticism with regard to knowing the Real? If perception does not ultimately
yield some immediate apprehension of the Real, although it provides occasion for the refort
nation​ ​of​ ​habit,​ ​must​ ​the​ ​hope​ ​of​ ​real​ ​certainty​ ​in​ ​our​ ​inquiries​ ​be​ ​relinquished?

In part Peirce has already answered the question by redefining the Real in terms of the
final opinion of inquirers in the long run. In any of our judgments certainty is not given in the
here and now by some Kantian application of categories but rather is relegated to the
inductively confirming test of time among an intersubjective community of inquirers.
Furthermore, what is known of the Real is not some set of spatio-temporal characteristics
defining existent objects but rather the overarching complex of laws and patterns which
regulate the things of the world. Peirce said that the Real in its totality is in essence
equivalent to the final opinion reached in inquiry. And what is opinion for Peirce if not belief,
or more pertinently, habit. So, in inquiring into the Real, we have knowledge of it to the
extent​ ​that​ ​our​ ​habits​ ​of​ ​inference​ ​exemplify​ ​the​ ​laws​ ​or​ ​habits​ ​of​ ​the​ ​Real.

In this sense the charge of skepticism is somewhat misplaced because it assumes a


conception of the Real which Peirce does not share, namely the Kantian sense of it as
objective and existentially there, hopefully amenable to an intimate correspondence in our
acts of judgment. Though it be true that perceptual judgment is inferential and thus liable to
error, Peirce does not believe this should lead us to a skepticism. If we find that one of our
judgments

was mistaken for example, that itself is a judgment made possible only by other perceptual

43
judgments. That is to say that the elements of our cognitive life are not isolated but rather
interact in a complex web of association that compares, supports, and criticizes. Mental life is
a continuum of inferential activity which, to return to Peirce's simile, is as strong, not as its
weakest link, but to the extent that it is formed by many tightly and closely bound fibers. In
this sense we can say that perceptual judgment is on the whole reliable, if in fact we find that
in​ ​any​ ​particular​ ​case​ ​we​ ​may​ ​have​ ​misjudged.

His analysis of perception can serve as occasion here to illustrate two methodological
themes which run throughout the broad course of his thought. The first of these is his
Fallibilism, which I have already alluded to. Basically, it is a stance assumed in inquiry that
acknowledges the finite limitations of the human knower and which accepts the experience of
error not as an obstacle to knowledge but as a necessary rung in the ladder of our climb
toward it. Our inferences are indeed fallible, but what redeems them is that they are also
self-correcting and thus we can inquire confidently in the face of error knowing that truth is
not thereby sacrificed. Indeed, to hold expectations of Cartesian certainty in our cognitive
acts would be to expect some sort of machine precision from us. As Hookway states,
"Cartesians court skepticism because they make an unreasonable demand for absolute
certainty;​ ​the​ ​moderate​ ​fallibilist​ ​lowers​ ​his​ ​sights,​ ​and​ ​settles,​ ​reasonably,​ ​for​ ​less."21

The other theme is what Peirce called his Critical Common-Sensism. It is the attitude
which accepts as necessary in inquiry the whole spectrum of our beliefs, opinions,
perceptions, ideas, and prejudices. Philosophy must begin en medias res and work itself out,
rather than from some ideal point of first beginnings. Tempered with Fallibilism this attitude
supposes that our inquisitive abilities do not lead us into wholesale error on every front. It
recognizes that perdurable patterns of belief reflect successful attempts at the truth, while
remaining​ ​always​ ​subject​ ​to​ ​revision​ ​in​ ​the​ ​course​ ​of​ ​experience.​ ​As​ ​Peirce​ ​said,

[Upon] innumerable questions, we have already reached the final opinion. How do
we know that? Do we fancy ourselves infallible? Not at all; but throwing off as
probably erroneous a thousandth or even a hundredth of all the beliefs established
beyond present doubt, there must remain a vast multitude in which the final opinion
has been reached. Every directory, guide book, dictionary, history, and work of
science​ ​is​ ​crammed​ ​with​ ​such​ ​facts​ ​(8.43).

21
​ ​Hookway,​ ​p.​ ​73.

44
It is clear from this that Peirce is more than sanguine about our ability to reach the truth. This
optimism is grounded in "his profound insight into the nature of inquiry, one which has the
effect of reconciling the paradox that error is a necessary element in the attainment of truth
with the common-sense position that truth is always there to be found in any field in which
we​ ​know​ ​how​ ​to​ ​inquire.​"22

To paraphrase Socrates, it is not just a matter of inquiring, but of inquiring well. In the
formation of a new belief it is true that inference is motivated only by the removal of the
irritation of doubt, but it operates as well under the exigency that any new belief should
answer to experience more potently and efficiently than those that went before, in short that
belief should be perdurable, and thus that our inferences should be guided by a method which
checks the vagaries of idiosyncrasy. Inquiring well, then, entails the method of science as
Peirce outlined, and the hypothesis of reality which it implies. Although his analysis of
inference showed that many of the leading premises which guide inference are a-critical or
even wholly unconscious, it is the critical attention paid to the pointedly conscious activity of
inference that we should cultivate and develop if inquiry is to achieve its deepest aims. The
deliberate control of reason is the catalyzing edge of inquiry as it moves toward the final
opinion.

But what does it mean to be critical? If the truly ampliative component of inquiry lies in
abductive inference, what tools or methods should be used in its control? At this point we
could move directly to a discussion of the pragmatic maxim, for it is precisely this use that
Peirce envisaged in its conception, that is, as a tool in the admissibility of hypotheses.
However a full appreciation of the potency of the maxim would be lost without first
understanding Peirce's conception of the sign. The process of inference describes the activity
of thought, the structure within which it moves, but an account of inquiry would remain
incomplete if the material which comprises thought were not considered. As Peirce
concluded in "Some Consequences of Four Incapacities", we have no power of thinking
without signs. "Now thought is of the nature of a sign," thus, "if we can find out the right
method of thinking -- the right method of transforming signs -- the truth can be nothing more
than​ ​the​ ​result​ ​to​ ​which​ ​.​ ​.​ ​.​ ​this​ ​method​ ​would​ ​ultimately​ ​lead​ ​us"​ ​(5.553).

22
​ ​Hookway,​ ​p.​ ​107.

45
A discussion of semiotic is important because sign activity is the means by which the
community communicates and thus is able to develop and use the results of inquiry. It will
also provide the context for an understanding of Peirce's categories and of his conception of
the normative sciences of ethics and aesthetics which will be critical in understanding the
later reorientation he gives to the pragmatic maxim. Finally, according to Apel, a treatment of
semiotic will complete and make comprehensible Peirce's transformation of Kant's
epistemology, clearing the way for an evaluation of his later pragmatism in terms of the
transcendental​ ​concerns​ ​he​ ​took​ ​from​ ​Kant.

Chapter​ ​Two

46
Two​ ​Semiotic

There is perhaps no other topic about which Peirce wrote more pervasively and deeply,
and that has had as much subsequent influence, than his writings on semiotic. Its prominent
role in his thinking is due in large measure to the fact that many areas of philosophical
investigation - epistemology, as well as metaphysics and ontology, are explicable in terms of
sign activity. As will become clear later, Peirce's understanding of the world as pansemiotic,
that is, that the constituents of the Real, from the knowing inquirer to the universe which is
known, are all terms in a fundamental process of sign activity, is what allows him to obviate
the many pitfalls of modern Cartesianism, and is what I esteem as his most important and
invigorating​ ​contribution​ ​to​ ​philosophy.

For various reasons his contribution fell on mostly deaf ears, and since then it has
languished outside the community of inquirers. It is only in recent years, in the current
renaissance of his thought, that it has begun to receive it's due. As will be seen, a
distinguishing feature of Peirce's semiotic is that it is triadic in nature. For a sign to be
operative at all its activity must be characterized by three irreducible elements (sign, object,
and interpretant). In introducing his thinking on signs however I want to anticipate the power
of his theory and throw it into relief by first briefly examining the conception of semiotic
which has held sway in much contemporary thought, namely that of Ferdinand de Saussure. I
do this because Saussure's conception of semiotic is dyadic in nature and thus provides a foil
for introducing Peirce's triadic structure. In addition it provides an epistemological model
necessary in understanding the relativistic thrust of contemporary theory. In so doing we can
more​ ​easily​ ​evaluate​ ​Peirce's​ ​conception​ ​as​ ​an​ ​answer​ ​to​ ​contemporary​ ​dilemmas.

A further important preliminary in understanding Peirce's notion of triadicity is the


development of his categories. The structure of his semiotic is based on this conception and
being familiar with its terms will help immeasurably in its clarification. I intend for my
discussion to be brief, aimed not at an exhaustive analysis and justification of his derivations,
but​ ​at​ ​basic​ ​definitions​ ​that​ ​will​ ​aid​ ​in​ ​the​ ​use​ ​of​ ​its​ ​terms​ ​in​ ​understanding​ ​his​ ​semiotic.

Saussure​ ​and​ ​Signification

47
Like Peirce, Saussure saw that much could be explained by understanding language as a
system of signs. In asking about the meaning of a sign Saussure said that a distinction must
be made, one between language and speech, or la langue and la parole. Language as a system
is what he called langue, and the particular manifestations of it in speech, parole. As a
practicing linguist what was of practical importance to him was parole, for this provides the
data of investigation. But as a philosopher what was of abiding interest was langue itself, the
system which gives speech meaning in any particular context. Langue is an institutional set of
impersonal rules, a system of conventions which we assimilate in the process of learning to
communicate. Without its rules it is not possible to speak of the meaning of any particular
speech act. Thus, as a philosopher of language Saussure is interested not in what a sign
means,​ ​but​ ​how​ ​in​ ​fact​ ​it​ ​is​ ​able​ ​to​ ​mean,​ ​with​ ​the​ ​structure​ ​of​ ​the​ ​system​ ​of​ ​signs.

In​ ​elucidating​ ​this​ ​structure​ ​Saussure​ ​began​ ​with​ ​the​ ​structure​ ​of​ ​the​ ​sign​ ​itself,

I call the combination of a concept and a sound-image a sign, but in current usage
the term generally designates only a sound-image . . . . I propose to retain the word
sign to designate the whole and to replace concept and sound respectively by
signified and signifier; the last two terms have the advantage of indicating the
opposition that separates them from each other and from the whole of which they
are​ ​parts​ ​.​ ​.​ ​.23

Signifier and signified, sign and object, represent the two terms of a speech act, and are
characterized by being completely arbitrary in their relation. There is nothing more adhering
between any sign and the object for which is stands than mere convention. And yet, though
the sound we utter to mean 'cat' could just as easily be used to refer to a 'dog', we are not at
liberty as individuals to do so. For Saussure, the essentially arbitrary nature of actual speech
is fixed by the particular structure of a system of signs in any given language. Though the
relation between signifier and signified may change over time this is a function of cultural
evolution​ ​and​ ​not​ ​individual​ ​whim.

Again, it is langue that, as a system, provides the basis for signs having meaning in
everyday speech. The positive value that a sign has is a function of its position within the
system of all signs such that its identity is constituted in its difference from other signs, and
by nothing else. In formulating the mechanics of signification in this way Saussure disavows
for the study of signs any recourse to historical analysis or metaphysical speculation in the

23
​ ​Saussure,​ ​Ferdinand​ ​de,​ ​Course​ ​in​ ​General​ ​Linguistics,​ ​(Paris:​ ​Payot,​ ​1960).​ ​p.​ ​14

48
evaluation of a sign's meaning. This is given in his distinction between the diachronic and
synchronic branches of linguistic analysis: the former investigates historically, the latter
structurally. Synchronic analysis brackets temporal elements in the question of meaning,
making essential only a sign's position and difference within a system. It was this distinction
which laid the theoretical groundwork for Structuralism in this century. As Sheriff has
summed​ ​up​ ​the​ ​consequences,

The prevailing orthodoxy that has emerged is that in language there are only
differences without positive terms. A sign has no meaning in itself but has
meaning (only in relation to other signs. Consequently, meaning depends on the
system,​ ​on​ ​the​ ​oppositions​ ​and​ ​contrasts​ ​within​ ​the​ ​system.24

It is the terms of Saussure's analysis which gives thrust to contemporary thought. In his
critique of the metaphysics of meaning Jacques Derrida appropriates Saussure's structure and
carries it to its logical conclusion. Meaning is not some positive given to be discovered, but is
the product of a system of differences which creates it. What we grasp in meaning is not the
thing signified but "the space, the gap, where the thing signified is lost in order to be
signified"25. It is this sense which characterizes the ultimately deferring character of
signification and thus the fact that meaning confers not the presence of the signified but its
trace.

This gap between signifier and signified is the bugbear of any thought which attempts to
grapple with contemporary relativism. The resilience of the host of current deconstructive
discourses stems partly from the perception that the only alternative is to return to the
mystifications of Enlightenment conceptions of meaning. But in addition, as Peirce's analysis
of inference showed, semiotic conceptions of language and cognition have much more
explanatory power than traditional epistemologies. We read Derrida and we can't help but to
admire (if sometimes secretly) the power of his analyses, but we are also left discomfited, and
feel that something is awry in all this as we examine our profoundest insights into the nature
of​ ​meaning.

But this is no wonder if we persist in accepting the conception of language as a system

24
​Sheriff, John K., The Fate of Meaning: ​Charles S. Peirce, Structuralism, and Literature​, (Princeton: Princeton University
Press,​ ​1989),​ ​p.​ ​11.
25
​ ​Ibid.​ ​p.​ ​37

49
of dyadic signs. If there is no logical, natural or otherwise determined cohesion between the
parts of such a system, if the basis for relation is merely a matter of convention, then meaning
will remain ultimately relative, being determined not by a signifier but by a blind system of
differences. The improvement Peirce makes here is found in his conception of signs as
essentially triadic in nature. Whatever this third element is it will confer on interpretation the
constraint​ ​of​ ​something​ ​more​ ​enduring​ ​than​ ​mere​ ​convention​ ​or​ ​whim.

The​ ​Origin​ ​of​ ​Peirce's​ ​Semiotic​ ​in​ ​the​ ​Derivation​ ​of​ ​the​ ​Categories

The tripartite structure of semiosis has its logical genesis in the derivation of the formal
categories of thought, to which I now turn. Peirce's principle discussion of the categories is
given in the article, "On a New List of Categories," published in the Proceedings of the
American Academy of Arts and Sciences in 1867.26 The title is a clear reference to his
continuing engagement with and transformation of Kant. In line with his general reappraisal
of the cognitive process, he proceeds in this paper to derive a list of categories pertinent to his
new conception. In philosophical discussions in general the notion of categories refers to
conditions of intelligibility. Drawing such a list is an attempt to specify the broadest
conditions under which all things may be classified, experienced, and known. As Hausman
says, "Such conceptions may be regarded as the classes or types of things into which things
that are and can be known can be divided."27 Thus Peirce aims in this paper to give more
formal structure and depth to the process of inquiry by specifying the general conditions
under​ ​which​ ​conceptions​ ​may​ ​be​ ​formed.

As categories essentially of thought Peirce is concerned with propositions, insofar as


propositions are the manifestation of knowledge, or are the result, to speak in Kantian terms,
of bringing the manifold of sensuous impressions to a unity. The question is, "What
conceptions​ ​are​ ​necessarily​ ​involved​ ​in​ ​the​ ​formation​ ​and​ ​intelligibility​ ​of​ ​any​ ​proposition?"

This​ ​question​ ​is​ ​answered​ ​succinctly​ ​in​ ​the​ ​following​ ​passage,

Three conceptions are perpetually turning up at every point in every theory of


logic. . . They are conceptions so very broad and consequently indefinite that they
are hard to seize and may be easily overlooked. I call them the conceptions of
First, Second, Third. First is the conception of being or existing independent of
anything else. Second is the conception of reaction with something else. Third is

26
​ ​Vol.​ ​7​ ​(May​ ​1867)​ ​pp.​ ​287-98.
27
​H​ ausman​ ​p.​ ​94.

50
the conception of mediation, whereby a first and second are brought into relation.
(6.32)

Any proposition which brings a manifold to unity exhibits what Peirce calls Firstness,
Secondness,​ ​and​ ​Thirdness.

I will speak more concretely in a moment about what Peirce means by each of these, but
first I want to briefly discuss their derivation. The categories can be seen as the result of two
types of investigation: phenomenological, and logical. His phenomenological analyses are
long and very detailed but succeed in Peirce's judgment in specifying the three conditions
noted above in any experience examined. More incisive in my view is his logical analysis. It
derives from his varied and diverse writings on the logic of relations and concerns an analysis
of​ ​propositional​ ​structure.​ 28

For Peirce what cognition achieves in the formation of a proposition is the unification of
monadic characters with the elements of experience. That is to say that things are brought
together or unified by being predicated. Taking "Socrates is mortal" as an example, we can
easily see the subject-predicate relation. What interests Peirce though is the description of the
fundamental conditions under which such a unification is possible. The expression ". . . is
mortal" is for Peirce an 'unsaturated' expression. It remains to be filled by a proper name so as
to complete or 'saturate' it.29 Whenever a proposition is formed the sites of its 'attachment' or
what Peirce terms 'valence' must be filled. The example above has one valence but other
expressions have more, for example, . . . married . . . ", or ". . . gives . . . to . . . ".
Understanding the conditions of unification then entails the "classification of incomplete
expressions​ ​according​ ​to​ ​their​ ​valence".30

From the statement of his categories above it is clear that Peirce conceives of three

28
​I am indebted to Hookway's analysis for my account of Peirce's understanding of logic and the categories. For a clear and
detailed analysis of the phenomenological derivation see Hookway, ​Peirce. (London: Routledge & Kegan Paul plc, 1985),
pp.​ ​101-12.​ ​Passages​ ​in​ ​which​ ​Peirce​ ​discusses​ ​the​ ​derivation​ ​are​ ​at​ ​.284-353​ ​and​ ​5.4-65.

29
Both Peirce and Gottlob Frege employed this terminology in their contemporaneous work on the structure of predication.
For references to Frege's work see, Gottlob Frege, The Basic Laws of Arithmetic Exposition of the System. ​trans. and ed. by
Montgomery Furth (Berkley: University of California Press, 1964). #1, p. 34. See also, "Function and Concept", ​Collected
Papers on Mathematics, Logic, and Philosophy​, translated by Max Black, et. al., edited by Brian McGuinness (Oxford: Basil
Blackwell,​ ​1984),​ ​pp.​ ​140-l.

30
​ ​Hookway​ ​p.​ ​87.

51
types of incomplete expression -- ones with valences of one, two, or three. Expressions with
valences of four, five and higher certainly exist but logically they are reducible to
combinations of three-termed expressions. Thus does Peirce conceive of the categorial
structure of all propositions, that upon analysis they manifest monadic, dyadic, and triadic
predicates. The general concepts or categories expressed in this formulation are Firstness,
Secondness,​ ​and​ ​Thirdness.

To return then to a description of the categories. Firstness is the most difficult of all the
categories to describe because, though it is a logical element of any experience, it is not one
we are consciously aware of. To reflect on and describe experiences of Firstness is already to
be engaged in what we will see to be the mediating character of Thirdness. In essence,
Firstness refers to the simple positive quality of something, as it is without relation to
anything else. He says, "Such a consciousness might be just an odour, say a smell of attar; or
it might be one infinite dead ache; it might be the hearing of a piercing eternal whistle. In
short, any simple and positive quality of feeling would be something which our description
fits that it is such that it is quite regardless of anything else" (5.44). Peirce is stressing here
the​ ​monadic​ ​quality​ ​of​ ​all​ ​the​ ​possible​ ​elements​ ​in​ ​experience.

To confound things he say that that which Firstness exemplifies is not anything existent,
but is rather mere possibility. If this 'positive quality' is realized then it no longer stands
without relation. To re-emphasize, what Peirce means to describe by Firstness is a logical, not
an experiential component of experience, which thus makes description of it a difficult task.
Maddeningly, Peirce says that "... every description of it must be false to it" (1.357). It is
approachable​ ​only​ ​to​ ​the​ ​extent​ ​which​ ​we​ ​precind​ ​from​ ​experience.

However, his description of it as standing without relation leads to a consideration of his


second category. By Secondness Peirce refers to the reactive, dyadic element of experience,
of things standing in brute relation to one another. It is the category of existence, connoting
such things as reaction, effort, resistance, shock, and surprise. Secondness describes the
conjunction of two things with one another, their relation constituting a sheer brute fact. Our
awareness of it is given in experiences of sudden alteration, such as the turning on of a light
in a dark room, or the jolt of an electric shock when you were sure that the outlet was
grounded. Secondness is more amenable to description because it is a real element of our
experience, but it still remains just outside of a fully cognitive grasp. Events of secondness

52
happen suddenly and then are gone. Recalling them and reflecting on them is an activity
which​ ​the​ ​category​ ​of​ ​Thirdness​ ​describes.

Thirdness is the conception of mediation, the bringing together of two things, not
existentially as in Secondness, but intelligibly. The brute relation of two things in the
universe is sheer happenstance, an occurrence classifiable as nothing more than fact.
Secondness in this sense is descriptive of nothing more than a present existential state of
affairs. Thirdness however is concerned with the future. It is descriptive of relations which
predict patterns of future behavior, or as Peirce would say, of "would-be's'. In short,
Thirdness is the category of law, of regularity, habit and generality. As Gallie concisely
states,

An experience counts as First in virtue of what it immediately is; as Second in


virtue of what it immediately or actually does to other things; as Third in virtue of
its "would-be" and 'would-do' -- its capacity to influence any of a describable set
of later experiences by developing . . . the "would-be's' and "would-do's' of certain
experiences​ ​which​ ​preceded​ ​it.31

It is Thirdness which most accurately describes the inferential habits of mental life. In
this conception of its mediating role we have the most potent exemplar of it in the activity of
a sign as it mediates between an object and its interpretant. It is here that we can most
fruitfully enter a discussion of Peirce's semiotic, through the course of which I hope to
deepen​ ​and​ ​expand​ ​the​ ​insights​ ​of​ ​his​ ​categorial​ ​analysis.

The​ ​Triadic​ ​Theory​ ​of​ ​Signs​ ​and​ ​the​ ​Need​ ​for​ ​Interpretative​ ​Control

The​ ​significance​ ​of​ ​signs​ ​for​ ​Peirce​ ​is​ ​given​ ​in​ ​the​ ​following​ ​remark:

Know that from the day when at the age of 12 or 13 I took up, in my elder brother's room a copy
of Whatley's 'Logic,' and asked him what Logic was, and getting some simple answer, flung
myself on the floor and buried myself in it, it has never been in my power to study anything, --
mathematics, ethics, metaphysics, gravitation, thermo-dynamics, optics, chemistry, comparative
anatomy, astronomy, psychology, phonetics, economics, the history of science, whist, men and
women,​ ​wine,​ ​meteorology,​ ​except​ ​as​ ​a​ ​study​ ​of​ ​semiotic​ ​(SS​ ​85-6).32

31
​ ​Gallie​ ​p.​ ​200.

32
​Charles S. Hardwick, ed., ​Semiotic and Significs: The Correspondence between Charles S. Peirce and Victoria Lady
Welby​,​ ​(Bloomington:​ ​Indiana​ ​University​ ​Press,​ ​1977).

53
It seems not surprising then that Ransdell has estimated that a good 90 per cent of his
writings concern semiotic directly.33 Even if this errs on the side of exaggeration it is clear
that the great majority of Peirce's thought is formulated in terms of the heuristic conceptions
of​ ​sign​ ​activity.

What can be most generally said about a sign is that it is anything that tells us about
something other than itself, or more specifically, perhaps, that it is anything which makes
reference to or indicates something other than itself. To use the phrase "tells us' is to assume
that all signs are operative only for a human intelligence, and though Peirce often speaks in
these terms, given that his concern is with human inquiry and knowledge, the terms he will
use to define sign activity denote a much broader range of application, including phenomena
in the animal and perhaps even vegetable worlds. It is his concern with the autonomy of logic
and his critique of Cartesianism which serve him in constraining his analysis of signs from
being​ ​too​ ​anthropomorphic.

The 'something other' to which a sign refers Peirce terms the object of the sign. It is that
which we normally say a sign means. Sign and object are the signifier and signified in
Saussure's analysis and constitute the traditionally conceived dyad of signhood. This relation
is elemental to sign structure, and Peirce spends no small amount of time in analyzing it, but
there is something lacking, something which has the power to explain how a sign refers. Let
us​ ​refer​ ​then​ ​to​ ​a​ ​few​ ​of​ ​Peirce's​ ​many​ ​formulations,

A Sign, or Representamen, is a First which stands in such a genuine triadic relation


to a Second, called its Object, as to be capable of determining a Third, called its
Interpretant, to assume the same triadic relation to its Object in which it stands
itself​ ​to​ ​the​ ​same​ ​Object​ ​(2.274).
Grant, then, that every thought is a sign. Now the essential nature of a sign is that
it mediates between its Object, which is supposed to determine it and to be, in
some sense, the cause of it, and its Meaning. . . . the object and the interpretant
being the two correlates of every sign. . . . the object is the antecedent, the
interpretant​ ​the​ ​consequent​ ​of​ ​the​ ​sign.34

33
​ ​Ransdell,​ ​J.​ ​M.​ ​"Some​ ​Leading​ ​Ideas​ ​of​ ​Peirce's​ ​Semiotic",​ ​Semiotica​ ​(1977)​ ​vol.​ ​19,​ ​p.​ ​153.

34
​ ichard S. Robin, ​Annotated Catalogue of Charles Sanders Peirce​, (Ameherst: University of Massachusetts Press, 1967),
R
318,​ ​pp.​ ​323-32.

54
In addition to the object, the other significant correlate of a sign is what Peirce calls the
'interpretant'. Though Peirce often describes semiosis in terms of its use by an intelligent
inquiring mind, he makes this coinage to distinguish his conception from such
anthropomorphic notions as 'interpretation' or 'interpreter'. As stated in the first description
above, the interpretant denotes the creation of a further sign which stands in the same relation
to the object as does the original sign. The power of Peirce's conception of semiosis derives
in part from a terminology which gives explanatory scope not only to intelligent inquiry but
to extra-mental processes as well, thus helping to bridge the Cartesian gulf between humans
as​ ​knowers​ ​and​ ​the​ ​world​ ​that​ ​is​ ​known.

The elements of the sign triad then are sign, object, and interpretant and constitute the
necessary conditions for sign activity. An example will illustrate how Peirce understands
their operation. My roommate and I are sitting in the living room reading when suddenly I
hear a knock at the door and I turn my head in the direction of the door. My roommate, who
is deaf, sees the sudden turning of my head toward the door and takes this to mean that
something, probably a knock, directed my attention there. The original sign in this example is
the turning of my head, its object being the knocking at the door. The interpretant is that
which was created or determined by this dyadic relation, namely the turning of my
roommates' head toward the door. The interpretant is the meaning of the original sign, and is
itself a new sign which stands in the same relationship to the object as the original sign. Its
relation to the object is mediated by the original sign. In a similar way the interpretant, or the
new sign, can mediate between the object and a further new interpretant. This interpretant
might be another friend who was in the room listening to music with headphones, and thus
was not determined by the original knock. Being so positioned as to see, only my roommate
and not me, the turning of my roommates' head would act in the same sign-object way to
determine a new interpretant, perhaps a thought this time that attention in the living room has
been​ ​drawn​ ​to​ ​something.

What is immediately characteristic of this view of sign activity is its web-like nature.
Signs are not wholly and sufficiently connected to objects in a discrete series of facts about
the world, but are rather interconnected. They form a triadic lattice which is open-ended, the
operative force of which is given in the ability of every generated interpretant to stand itself
as a further sign in what appears to be an endless semiotic process. It should be noted here

55
that sign, object, and interpretant are fluid terms and denote the various elements of the
process depending on the point of view taken. For example the original sign which
determines an interpretant becomes the object when the generated interpretant acts as a sign
in​ ​determining​ ​a​ ​new​ ​interpretant.

But it still may be asked why this triadic structure is necessary if signs are to generate
meaning. Referring to Saussure's definition we can ask why the above example cannot be
explained in terms of two dyadic relations. Peirce describes the situation in the following
statement,

An event A, may, by brute force, produce an event, B; and then the event, B may in
its turn produce a third event, C. The fact that the event, C, is about to be produced
by B has no influence at all upon the production of B by A. It is impossible that it
should, since the action of B in producing C is a contingent future event at the time
B is produced. Such is dyadic action, which is so called because each step of it
concerns​ ​a​ ​pair​ ​of​ ​objects​ ​(5.472).

The knock at the door occasioned the turning of my head. The turning of my head occasioned
the turning of my roommates' head in the same direction. The question is how these two
events can be seen to be related rather than distinct. Referring to his statement above,
triadicity is involved when A produces B as a means to the production of C. If B is likely to
produce​ ​C​ ​then​ ​A​ ​will​ ​produce​ ​B,​ ​but​ ​otherwise​ ​it​ ​will​ ​not.

This seems a strange situation indeed, as if the knocking at the door were meant to
induce the turning of my roommates' head. It seems undeniable though that something like
purpose informs his understanding of the sign relation. In showing how Peirce explains this
connection Hookway draws a helpful distinction. Referring to two apparently dyadic relations
he​ ​says​ ​that​ ​"there​ ​seems​ ​to​ ​be​ ​a​ ​distinction​ ​between​ ​the​ ​following,

(l)​ ​Event​ ​E​ ​causes​ ​us​ ​to​ ​believe​ ​that​ ​P.

(2)​ ​We​ ​take​ ​E​ ​as​ ​evidence​ ​for,​ ​or​ ​as​ ​a​ ​sign​ ​of,​ ​the​ ​fact​ ​that​ ​P."35

Hookway notes that in the first instance we need have no opinion or familiarity with E in
coming to our belief about P whereas with the second case there must be some engagement or
evaluation of P. To put this in terms of another example, smoke seen in the distance causes us
to believe that there is a fire. First the fire causes the smoke and then either case (l) or case

35
​ ​Hookway​ ​p.​ ​123.

56
(2) above applies. Case (l) would be an instance of a particular class of signs whose function
Peirce calls indexical. The sign is an index of its object because there is a direct causal
relationship between the two. The fire causes the smoke. To say that the apprehension of the
smoke and the ensuing belief is similarly indexical is insufficiently explanatory. The
interpretant belief is not produced indexically, in a causal manner, but rather triadically. It is
the interpretation of the sign in terms of its relationship to its object that makes of signs
vehicles​ ​of​ ​meaning.

That fire causes smoke is simply a brute dyadic fact, one among many in the world. But
to say that smoke signifies fire is to describe the triadicity that Peirce termed Thirdness.
Thirdness is the category of meaning, the movement beyond singular fact to repeatability and
regularity. This is not to diminish the relevance of the dyadic relation in sign activity. As
Peirce​ ​describes​ ​below,​ ​it​ ​plays​ ​a​ ​vital​ ​role​ ​in​ ​the​ ​constraint​ ​of​ ​interpretation,

Now a sign has, as such, three references: first, it is a sign to some thought which
interprets it; second, it is a sign for some object to which in that thought it is
equivalent; third, it is a sign, in some respect or quality, which brings it into
connection​ ​with​ ​its​ ​object​ ​(5.283).

Peirce says here that a sign is referent to some respect or quality. He refines this by
elsewhere​ ​saying:
A sign, or representamen, is something which stands to somebody for something
in some respect or capacity. The sign stands for something, its object. It stands for
that object, not in all respects, but in reference to a sort of idea, which I have
sometimes​ ​called​ ​the​ ​ground​ ​of​ ​the​ ​representamen.​ ​(2.228)

The ground he speaks of here is what provides the 'intentional' thread between the three
relata. It refers to the nexus of beliefs and experiences which guide our inferences with regard
to particular things like smoke, or anything else that acts as a sign. Our inductive familiarity
with fire and smoke is the ground which constrains interpretation and thus the creation of the
usual interpretant "Fire!" Were we to revise our beliefs regarding the effects of fire, then
smoke would no longer have the power to induce the interpretants that it does. As Hookway
says, "This provides a particularly clear example of the triadic production of interpretants; it
is a case where our interpretation of the sign rests upon conscious reflection and upon beliefs
about​ ​the​ ​relations​ ​between​ ​sign​ ​and​ ​signified​ ​which​ ​ground​ ​the​ ​sign's​ ​function."36

36
​ ​Ibid.​ ​p.​ ​124.

57
The triadic nature of sign activity though is often masked in ordinary usage. In saying
for instance that "X means Y" or that "A interprets B", one of the elements is abstracted and
what is uncritically apprehended as the functioning of a sign is a dyad. However, regarded in
their abstracted form it is possible to speak of different types of signs. Based on his categorial
analysis Peirce was able to distinguish sixty-six classes of signs. His first division, though, is
into three main classes and it is these which provide the schematic form for all which follow.
The three types of sign are Icon, Index, and Symbol. They are distinguished according to the
grounds which make them functional. To recall, a sign's ground is the relationship adhering
between​ ​it​ ​and​ ​its​ ​object,​ ​thus​ ​it​ ​is​ ​the​ ​sign/object​ ​relation​ ​which​ ​factors​ ​in​ ​the​ ​classification.

As I briefly mentioned in the example above, the presence of smoke is a sign that Peirce
would classify as an Index. Any such sign derives its semiotic potency from a direct causal
relationship to its object. It is bound in a dynamic and existential way to that which it
signifies, as in the relationship between a weather vane and the wind. The vane is an
indexical sign of the wind because of the causal relation between the two. The brute dyadic
associations of the category of Secondness are apparent here. A street-map on the other hand
is a sign which functions iconically. Its relation to its object is one of resemblance. Unlike an
indexical sign its object in fact cannot exist, yet its power to signify remains. A toy truck, a
blue print, a golf ball held against the night sky are all icons because they exploit a likeness
with the object they signify. Finally, a third kind of ground is one which designates signs as
symbols. A red stop light, for example, is a symbol, as are the written and spoken words of
most languages. Adhering in this sign/object relation is not resemblance, nor causation, but
mere convention, the existence of a general practice by which something is meant to indicate
something​ ​else.

By making this trichotomy Peirce does not mean to say that a sign has to be one or the
other of the three. To take the example of a weather vane, it may be molded in the shape of a
rooster and thus iconically signify a real rooster, or as it turns in the wind it indexically
signifies the direction of the wind, or finally as something antique it may signify a bygone
age and the various associations we have of that. The upshot of all these distinctions is finally
no more than analytical, to be clear about the elements that make up sign activity. The point
to be emphasized is that whether a sign be Iconic, Indexical, or Symbolic, it cannot function
unless it is interpreted. The smoke rising from a campfire is just a brute fact in the world.

58
defined by its ground, but meaningless unless it be capable of producing an interpretant, in
short​ ​of​ ​being​ ​interpreted.

What is missing in Saussure's analysis is an account of the interpretative element in


semiosis. Clearly he assumed someone interpreting signs, but it is only implied and says
nothing about the dynamic of interpretation itself. As opposed to Peirce, this view of sign
activity is a static one and unwittingly perpetuates a fundamentally Cartesian split between
knower and known, as if the knower were a hermetic entity manipulating signs. It is precisely
this view which fosters the relativistic apparatus of contemporary theory. If signs have only a
dyadic, purely conventional relationship to their objects then interpretation is completely cut
loose​ ​from​ ​any​ ​concern​ ​with​ ​proper​ ​inquiry,​ ​and​ ​thus​ ​with​ ​truth.

Initially Peirce's triadic analysis would seem to bode a fate similar to Saussure's. If
every interpretant can stand as a further, refined sign to an original object, then the process of
semiosis must be described as a system of signs with unlimited potential for interpretation.
For Peirce there is no single sign for a given object and no single interpretant of a given sign.
Instead, the triadic lattice of signification provides for a great many potential relations. In
such a way Peirce evades the mire into which the traditionally Kantian concern with
categories of judgment stumbles. Truth is not intuited in one fell swoop of judgment but
rather​ ​develops​ ​in​ ​the​ ​form​ ​of​ ​a​ ​sign​ ​in​ ​the​ ​revisionary​ ​process​ ​of​ ​inferential​ ​inquiry.

Peirce gives his nod here to what will later become the concerns of our post modern era,
yet reigns in the analysis by emphasizing the potential infinitude of the interpretative process.
Interpretation is potential on both practical and logical grounds. Practically, a discussion on
any given topic could go on endlessly, however the exigencies of life demand a certain
economy of inquiry and so we find ourselves making judgments compelled by practical
concerns. This is not to say though that inquiry is given over to blind willfulness. As Peirce's
analysis showed, we are still compelled by the model of the scientific method if our concern
is with a fruitful and lasting realization of inquiry's aims. The attitude that informs this
intersection of practical needs with inquiry's final aims is that of Fallibilism, the stance that
accepts the potential incompleteness of all sign interpretation, and thus the possible revision
of​ ​any​ ​belief​ ​subject​ ​to​ ​future​ ​experience.

Peirce's Fallibilism though is practical and methodological. It is meant to guide the


excesses of inquiry, not to frame it in a metaphysical view that denies it of ever reaching

59
truth. This leads us to consider the logical sense in which the inferential process is only
potential. The process of inquiry is guided by the hypothesis of the Real. The signs
interpreted in our inferences refer to real things or states of affairs, and though each sign is
capable of determining numerous interpretants and is thus always essentially developable,
there exists a definite array of interpretations which fills the valences of relation between the
sign and the Real it signifies. This "super" interpretant if you will is what Peirce termed the
"entire​ ​general​ ​intended​ ​interpretant"​ ​of​ ​a​ ​sign​ ​(5.179).

Though there is no measure by which we can be sure of having attained such a degree
of interpretative clarity, that we have in fact done so Peirce believes to be more than evident.
As I quoted earlier, " . . . throwing off as probably erroneous a thousandth or even a
hundredth of all the beliefs established beyond present doubt, there must remain a vast
multitude in which the final opinion has been reached. Every directory, guide book,
dictionary, history, and work of science is crammed with such facts" (8.43). To some degree
this must be the case, for if we were completely mistaken and all of our guiding principles
were completely false, our experience would be thoroughly frustrated in its engagement with
the world. However, sign interpretation always takes place within a particular context, thus
any single interpretation is restricted by perspective. When we say that a sign means such and
such a thing it is this restriction which keeps it open to future interpretative refinement. By
the "entire general intended interpretant" Peirce refers to the "very meaning" (5.427) of a sign
and refers not to one person's interpretative experience but to that of the community of
inquirers as a whole. To recall, meaning for Peirce is not some abstruse mental datum, but
rather given in the form of belief, as demonstrable in the habits of inference which guide
inquiry. As these habits become more stable and effective in their predictive powers, the
significative​ ​scope​ ​of​ ​the​ ​signs​ ​which​ ​determine​ ​them​ ​becomes​ ​more​ ​fully​ ​articulate.

The question emerges then, similar to that concerning method in the settlement of
opinion, about how to best proceed in bringing about the general intended interpretant. How,
as Gallie says, "do certain lines of interpretation of a given sign lead us with peculiar
directness and efficiency toward this result?"37 Taking 'water' as an example of a sign to be
interpreted we can begin to approach an answer by assessing the interpretative practices of
someone like a child and someone like a scientist. The interpretants which arise when a child

37
​ ​Gallie​ ​p.​ ​129.

60
considers water are ones such as "satisfying thirst," "cleaning the body," or "playing in a
swimming pool." It is a line of interpretation which addresses practically relevant concerns
and is in fact what characterizes the majority of even an adult's everyday inquiry. Certainly
these are possible interpretants of the sign 'water' but their scope, as is clear, is limited by the
concerns restricting their generation. Though the habits which guide these interpretations may
conform perfectly to logical standards, what is missing is deliberate control of and attention
to​ ​such​ ​standards.

Every sign is developable in terms of further interpretants which can refine it, so the
question is, out of all possible interpretants how does one discern those which are logically
commendable in achieving a sign's entire general intended interpretant? The interpretative
practices amenable to this concern are those of the scientist, a scientist being understood as
anyone who controls his or her inquiry in such a way that logical questions such as "Why?",
"How?", or "By what means and on what evidence?" are generated. That is to say that a
scientific inquirer engages in interpretation in such a way that logical consequences arise
which will either falsify or confirm the result. This deliberate, logical control of inquiry
occupied a great deal of Peirce's concern, for it is only through such control that the meaning
of​ ​a​ ​sign​ ​becomes​ ​progressively​ ​more​ ​clear.

However, what is the nature of this logical control? As we sa w earlier, the kind of
inferences which form the leading edge of the process of inquiry are abductive inferences, or
more familiarly, the formation of hypotheses. In light of the preceding discussion we can now
understand hypothesis formation in terms of interpretation, in the triadic generation of
interpretants. Since the concern is not to follow every line of interpretation for a given sign to
its exhaustive end, but rather to discern some strategy which will lead with efficiency to its
general intended interpretant, the question is how do we go about this? How does Peirce
envisage a logical control of inquiry such that by a certain economy of interpretation we
admit only certain hypotheses as amenable to the fruitful development and revelation of a
sign's "very meaning?" The answer is contained in the directive of the pragmatic maxim, to
which​ ​I​ ​now​ ​turn.

61
Chapter​ ​Three

Pragmatism

This chapter will concern itself mainly with Peirce's pragmatism. I will show how the
pragmatic maxim evolved as a consequence of his investigations into the logic of inquiry and
discuss how it put Peirce dangerously close to a nominalist position. It is here that James'
appropriation and popularization of it step over the nominalist line. The substance of what
follows from that point will be to determine the success Peirce had in reclaiming for
pragmatism the sense he intended for it in the context of his realist metaphysics. Before I
consider his pragmatism however I want to situate the results of the first two chapters in
terms​ ​of​ ​Peirce's​ ​revision​ ​of​ ​Kant's​ ​problematic.

Apel's​ ​Situation​ ​of​ ​Peirce​ ​in​ ​Kant's​ ​Project

First,​ ​our​ ​evaluation​ ​of​ ​Peirce​ ​must​ ​not​ ​be​ ​superficially​ ​misled​ ​by​ ​his​ ​critical​ ​approach
to​ ​Kant.​ ​Though​ ​he​ ​has​ ​discarded​ ​the​ ​Cartesian​ ​assumptions​ ​at​ ​the​ ​base​ ​of​ ​Kant's
epistemology,​ ​and​ ​begun​ ​to​ ​fashion​ ​something​ ​wholly​ ​different,​ ​Peirce's​ ​ultimate​ ​aims​ ​can
only​ ​be​ ​fully​ ​understood​ ​if​ ​they​ ​are​ ​seen​ ​as​ ​informed​ ​by​ ​Kant's​ ​transcendental​ ​concerns.

We have seen a hint of this in the hypothesis of reality. The manner in which Peirce
briefly argued for its justification is akin to the strategy Kant used in the deduction of his
categories. The argument has come to be known as the "transcendental argument," and
proceeds by claiming that a certain epistemological position must be true if some undeniable
capacity or fact is to be possible. For Kant it was the self-identifying capacity of the
"transcendental apperception," bringing the flux of experience to synthetic unity, which
argued for the truth of the claim that the objects of experience have spatio-temporal
substantiality​ ​and​ ​causality.

Similarly Peirce argues for the indispensable role that the hypothesis of reality plays in
the process of inquiry by showing that its assumption is necessary if the aims of inquiry are to
be met. Though he is critical of Kant's conception of mind Peirce shares with him the concern
that logic should ground philosophical investigation if its results not be tainted by the undue
influence of any of the empirical sciences. His hypothesis of reality reflects Kant's
characterization of the categories, that they correspond to nothing empirical, or to use terms

62
pertinent to Peirce's context, that the Real does not depend on the will or opinion of any
individual. They both assume an objective world to be known, and that any attempt to
understand how that knowledge is attained must proceed by a logic which gives inquiry a
scope​ ​unrestricted​ ​by​ ​the​ ​fashion​ ​of​ ​psychological​ ​theory.

But the hypothesis of reality is only an aspect of the logic of inquiry and I use it only as
an example to illustrate his general allegiance to Kant's concerns. Following Apel's reading I
want to anticipate the interpretative direction my analysis of Peirce will take with respect to
his​ ​transformation​ ​of​ ​Kant.

Peirce says, as I have mentioned, that "the lock upon the door of philosophy" is not,
following Kant, so much in discovering how synthetical judgments a priori are possible but
how synthetical reasoning itself is possible. Peirce's reformulation of Kant's question is
inspired by his analysis of inquiry. The Cartesian-based cognitive process Kant conceived of
as going on in the head is now spread out for Peirce in an inferential process among a
community of inquirers. The scientific method by which inquiry proceeds gives Peirce's
epistemology the "fallibilistic" character I have mentioned, that is, that beliefs are provisional
and are always open to revision in the light of further evidence. This takes the question of the
truth of our judgments, which for Kant was a matter to be resolved in the deduction of the
categories, and propels it into the future. For Peirce the elucidation of reality is not given in
piecemeal moments of judgment, but rather is manifested in the evolution of our habits of
belief, habits which change and adapt, and whose demonstration of "truth" can only be a
function​ ​of​ ​inquiry​ ​in​ ​the​ ​long​ ​run.

Thus his question how synthetic reasoning itself is possible. The very Kantian way of
asking it however is what leads Apel to see Peirce as appropriating the broad outlines of
Kant's project. If then we are to understand Peirce as engaged in some form of transcendental
philosophizing it should be possible to see in his "unlocking" of the door of philosophy some
manner of argument similar to that of Kant's. The terns however will be different. With a
view to securing the foundation of judgment formation, Kant was concerned with a deduction
of the objective validity of the categories. For Peirce, however, judgments do not have to be
true in order to be valid. Apel shows Peirce demonstrating their validity in terms of the
validity, in the long run, of the method by which they are acquired, namely inference or
induction. As Apel says, " . . . the problem of the transcendental deduction is transformed for

63
Peirce​ ​into​ ​the​ ​problem​ ​of​ ​the​ ​foundation​ ​of​ ​induction".38

Kant begins in the deduction with an analysis of transcendental apperception and the
synthetic unity of apperception. This undeniable capacity we possess, as I described above, is
what allows Kant to argue for the truth of his epistemological propositions. To follow Peirce
in this Kantian interpretation it will be necessary to identify some starting point in Peirce's
thought equivalent to this and determine what epistemological claim he is able to make in
light​ ​of​ ​it.

According​ ​to​ ​Apel​ ​this​ ​equivalent​ ​will​ ​be​ ​Peirce's​ ​semiotic​ ​and​ ​the​ ​pragmatic​ ​maxim​ ​which
lies​ ​at​ ​the​ ​center​ ​of​ ​it.​ ​Having​ ​examined​ ​his​ ​semiotic​ ​it​ ​is​ ​to​ ​an​ ​analysis​ ​of​ ​pragmatism​ ​that​ ​I
now​ ​turn.

The​ ​Pragmatic​ ​Maxim

It is in the paper "How to Make Our Ideas Clear" that Peirce introduces the pragmatic
maxim. This paper follows "The Fixation of Belief" and is aimed at refining the ideas about
belief he set forth there. To recall, belief is what Peirce called a habit. It is a rule or a
disposition to act, guiding the decisions we make according to the consequences of action
that we expect. Belief that fire burns, for instance, is a belief that manifests itself in the
disposition​ ​to​ ​keep​ ​our​ ​hands​ ​away​ ​from​ ​a​ ​flame.

Beliefs, though, are often vague, and have contributed in no small measure to
philosophical disputes throughout the tradition. How are the contents of a belief to be
understood, and how is one belief to be distinguished from another? The problem here
regards the question of meaning. What are its identifying marks? To this end Peirce proposed
a tool, a maxim of thought, "Consider what effects that might conceivably have practical
bearings we conceive the object of our conception to have. Then, our conception of these
effects is the whole of our conception of the object" (5.402). Contained within this rather
confusing statement is a definition of meaning. When we are disposed to act in a certain way
we either consciously or unconsciously anticipate the consequences that acting on a belief
would have if we were to perform it. What Peirce is saying is that meaning resides in nothing
more​ ​than​ ​this,​ ​in​ ​our​ ​conception​ ​of​ ​the​ ​effects.​ ​As​ ​he​ ​says,

38
​ ​Apel​ ​p.​ ​48.

64
The essence of belief is the establishment of a habit; and different beliefs are
distinguished by the different modes of action to which they give rise. If beliefs do
not differ in this respect, if they appease the same doubt by producing the same rule
of action, then no mere differences in the manner of consciousness of them can
make them different beliefs, any more than playing a tune in different keys is
playing​ ​different​ ​tunes.​ ​(5.398)

An analogy to music is made here which Hausman has interpreted in a way that
supports one of Peirce's claims we examined in Chapter One. There we saw Peirce insist that
the aims of inquiry cannot be achieved through idiosyncratic subjective preference. Here
Peirce likens belief to a tune or a melody. The particular order of tones which constitute a
melody distinguish it from other melodies and make it unique, as differing modes of action
distinguish one belief from another. Pursuing the analogy Hausman asks "what is the
analogue that a belief has in relation to a different key in which a single tune may be played?"
39
. He answers, quoting Peirce, the "manner of consciousness" in which the belief is held, the
various contexts or perspectives within which a belief may function. In music, though a
different key will create tones with different sounds, it is the order of the tones in a particular
sequence, their pattern, which remains constant through any number of keys and which, for
Peirce, is really pertinent in distinguishing one melody from another. The analogy confirms
the way Peirce understands the resolution of doubt. The belief which appeases doubt is not
determined by a particular perspective or by any subjective conditions but by "an objective
rule, a pattern that is not limited to specific, actual circumstances," just as a tune is not
limited​ ​by​ ​the​ ​particular​ ​tones​ ​of​ ​a​ ​specific​ ​key.

Though pragmatism came to define a movement, and then a pervasive philosophical


orientation, Peirce never intended it as more than a methodological tool in the course of
inquiry. It was these very "manners of consciousness" which contextualize belief that Peirce
had in mind in formulating the pragmatic maxim. If inquiry, as an engagement among a
community of inquirers, is an interactive affair where beliefs are compared and analyzed,
then some method of weeding out bias and perspective is needed if the settlement of opinion
is​ ​to​ ​be​ ​firm.​ ​Peirce​ ​defined​ ​this​ ​method​ ​in​ ​the​ ​pragmatic​ ​maxim.

39
​ ​Carl​ ​R.​ ​Hausman,​ ​Charles​ ​S.​ ​Peirce's​ ​Evolutionary​ ​Philosopohy​.​ ​(Cambridge:​ ​Cambridge​ ​University​ ​Press,​ ​1993),​ ​p.​ ​38.

65
As the name of the paper in which it is expounded states, it is a way of making our
ideas clear, of achieving what he called a "third degree of clearness." His critique of
Cartesianism continues here as he alludes to Descartes' definition of certainty, namely that we
should accept the truth of a proposition only if we clearly and distinctly perceive it as such in
our mind. The overly subjective nature of these two grades are clearly insufficient for the
philosophical vision Peirce is describing. Like the three discarded methods of fixing belief
(tenacity, authority, and the a priori), Descartes' definition of certainty ignores the realist
presupposition​ ​of​ ​inquiry​ ​and​ ​thus​ ​is​ ​inadequate​ ​to​ ​the​ ​full​ ​aims​ ​of​ ​inquiry.

On one level the third grade of clearness Peirce hopes to achieve is methodologically
different from Descartes' conception. Instead of assessing the status of a belief by some
determination of a state of mind, the pragmatic maxim instead directs us to consider belief in
terms of conceivable consequences. The criterion is not a private feeling but rather of the
nature​ ​of​ ​something​ ​public​ ​and​ ​observable.

On a second level there is a metaphysical difference. Whereas Descartes' definition


concerns certainty, and thus truth, Peirce's maxim pretends to no such thing. It is not a
criterion of truth, but one of meaning, or more precisely of the distinction of meaning. In this
distinction is revealed the truly revolutionary nature of Peirce's epistemological
transformation of Kant. His analysis of inquiry took Kant's critique of knowledge to the level
of a critique of meaning. That is to say, by relocating the work of cognition into the structural
framework of a process of inquiry he transformed the question of truth into one of meaning,
or of sense as such. The unifying role of judgment in Kant is transferred to the inferential
process of inquiry and resurfaces in the formation of belief. It is not judgments of fact but the
formation of a belief which is the new cognitive unit for Peirce, and in order to use them in
the course of inquiry it is not necessary that we know their truth value but only their limits.
Whereas for Kant the application of the categories was what aided judgment in its
determinations, for Peirce it is the pragmatic maxim, that rule by which beliefs, manifested in
our disposition to act, are discerned. More than anything it is the pragmatic maxim which
throws​ ​into​ ​relief​ ​Peirce's​ ​new​ ​vision.

I want to return to Peirce's formulation of the maxim and consider it in more detail. To
restate, "consider what effects, that might conceivably have practical bearings, we conceive
the object of our conception to have. Then, our conception of these effects is the whole of our

66
conception of the object." When considering the meaning of a conception, Peirce directs us to
regard the object of that conception, to that which the conception refers. Instead of fishing
through vague mental constructions Peirce asks us to consider the effects, or the practical
bearings, of the object of our conception. In conceiving of these we are looking for those
effects which would make a difference in our expectations, in our disposition to act. The
subject matter of the maxim here is the conception of a series of conditional actions that have
specifiable effects. It is in discerning these effects that the meaning of an idea becomes
intelligible. Meaning for Peirce is manifested in habits which differ from one another to the
extent​ ​that​ ​their​ ​conceivable​ ​effects​ ​in​ ​action​ ​differ.​ ​As​ ​Peirce​ ​more​ ​clearly​ ​stated​ ​elsewhere,

In order to ascertain the meaning of an intellectual conception one should consider


what practical consequences might conceivably result by necessity from the truth
of that conception; and the sum of these consequences will constitute the entire
meaning​ ​of​ ​the​ ​conception​ ​(5.9).

That statement was written twenty-five years after "How To Make Our Ideas Clear" as
Peirce struggled to reclaim the sense of the maxim he had intended for it. That the maxim got
loose from him and evolved into the form that it did is due in some part to the ambiguous
manner in which he talked about "effects" and "practical bearings", expressions which lend
themselves to a positivist or nominalist reading. In these early writings it seems that Peirce
talks about effects in terms of sense experience, the object of conception being reduced to or
equated with sensible effects. This would make of meaning not something general, which is
how he understood it, but particular and idiosyncratic. In later years he was self-critical of the
inadequacy of his initial formulation, but the damage was done. Most demonstrative of the
problem is the now infamous application of the maxim he gave to the meaning of the term
"hard."

Pragmatism​ ​and​ ​the​ ​Threat​ ​of​ ​Nominalism

By applying the pragmatic maxim to the conception of hardness we should discern


those effects which our expectations of it have and which thus distinguish it from other
qualities.​ ​The​ ​meaning​ ​of​ ​hardness,​ ​"what​ ​we​ ​mean​ ​by​ ​calling​ ​a​ ​thing​ ​hard"​ ​is

evidently that it will not be scratched by many other substances. The whole
conception of this quality, as of every other, then, lies in its conceived effects.
There is absolutely no difference between a hard thing and a soft thing so long as

67
they​ ​are​ ​not​ ​brought​ ​to​ ​the​ ​test"​ ​(5.403).

What Peirce is describing here is an actual test such that the meaning of hardness (when we
apply it to a diamond for example) would be all those consequences experienced through the
course of testing our expectations. It seems that he is conflating meaning with the sensible
effects​ ​borne​ ​of​ ​some​ ​testing​ ​procedure.

Peirce further probes the consequences of the maxim by considering a situation where a
diamond is crystallized in the midst of a soft cushion of cotton, and remains there until it is
finally​ ​burned​ ​up.

Would it be false to say that the diamond was soft? This seems a foolish question .
. . yet we may, in the present case, modify our question, and ask what prevents us
from saying that all hard bodies remain perfectly soft until they are touched, when
their hardness increases with the pressure until they are scratched . . . there would
be​ ​no​ ​falsity​ ​in​ ​such​ ​modes​ ​of​ ​speech​ ​(5.403).

In saying this Peirce brings himself dangerously close to a nominalism for which terms
have no generality and denote nothing more real than their convenience in aiding our
negotiation of reality. In light of this it is not difficult to see how James came to the
psychological interpretation he gave to Peirce's maxim, and we may ask how it is that Peirce
came to make such remarks given his critique of psychologism in his early papers. Perhaps it
was due in part to the influence of his nominalistically oriented friends in the Metaphysical
Club in Cambridge, Massachusetts in which he was involved at the time. Whatever the case
he clearly regretted the example and referred to it in the closing years of the century as he
reoriented his notion, now called 'pragmaticism', in the context of work done in the
intervening​ ​years​ ​on​ ​the​ ​metaphysics​ ​of​ ​evolution​ ​and​ ​the​ ​structure​ ​of​ ​the​ ​normative​ ​sciences.

At a later time Peirce gives an example which more clearly illustrates the parameters he
intended for the maxim, "The inkstand upon the table is heavy by which we only mean, that if
its support be removed it will fall to the ground. This may perhaps never happen to it at all --
and yet we say that it is really heavy all the time" (7.341). To refer to the diamond example it
is to say that we do not conceive its beginning to be hard only when we try to scratch it with
something. Peirce means to illustrate that when we employ the maxim, meaning is
determined not by some actual test, but by a real law manifested by the effects that would

68
result if we were to test it. Employment of the maxim is a thought experiment and turns on its
use​ ​of​ ​counterfactual​ ​conditionals,​ ​not​ ​actual​ ​tests,​ ​in​ ​explicating​ ​meaning.

Peirce opposes the predication of any quality such as hardness or softness as such to
their​ ​intellectual​ ​apprehension​ ​as​ ​a​ ​concept.​ ​As​ ​he​ ​says,

Intellectual concepts . . . essentially carry some implication concerning general


behavior . . . and so convey more, not merely than any feeling, but more, too, than
any existential fact, namely, the "would-acts," "would-dos" of habitual behavior;
and no agglomeration of actual happenings can ever completely fill the meaning of
a​ ​"would-be"​ ​(5.467).

Meaning is thus something general, not particular, and is revealed, in our use of the maxim,
not as a function of any particular act but of the habits which guide those acts. John Dewey
makes​ ​clear​ ​the​ ​distinction​ ​between​ ​Peirce​ ​and​ ​James​ ​when​ ​he​ ​says,

... one can say the he James enlarged the bearing of the principle by the
substitution of particular consequences for the general rule or method applicable to
future experience. But in another sense this substitution limited the application of
the principle, since it destroyed the importance attached by Peirce to the greatest
possible application of the rule, or the habit of conduct -- its extension to
universality. That is to say, William James was much more of a nominalist than
Peirce.40

More troubling at this point than his anomalous remarks about the diamond is the
implication of the maxin for the hypothesis of reality. In accepting the scientific method
Peirce's inquirer is committed to a notion of the real which, in guiding his inferences, is
ultimately unaffected by his particular idiosyncrasies, or by anything he may think. To recall,
in referring to the Real, Peirce speaks in terms of his logic of inquiry in which the Real is not
the world as it is constituted by the totality of existents, but rather the totality of laws which
govern it, laws which Peirce prefers to understand as manifested in habits. Habits of course
are beliefs or opinions and as such the Real, or "truth," is defined as the "final opinion of the
community of inquirers in the long run." The Real then is of the nature of a thought or a
belief, and the problem that arises here regards the independence of the Real. If the Real
which affects my inferences is nothing more than an agglomeration of other beliefs
developed thus far, then how can I be confident that my inquiry moves beyond merely finite

40
​John Dewey, "The Development of American Pragmatism", in ​Philosophy and Civilization (New York: Capricorn Books,
1963),​ ​p.​ ​14.

69
perspectives. Is the object represented by the final opinion, that is, the Real, independent of
that opinion? The difficulty deepens as we find Peirce remarking, "The object of the belief
exists, it is true, only because the belief exists, but this is not the same as to say that it begins
to exist first when the belief begins to exist" (7.340). Or is it? Peirce's statement to the
contrary does not convince. It would seem that the dilemma encountered in the diamond
example resurfaces here, although it is not as easily resolved, for it concerns not a particular
aspect​ ​of​ ​inquiry​ ​but​ ​the​ ​nature​ ​of​ ​the​ ​very​ ​presuppositions​ ​which​ ​ground​ ​it.

This application of the maxim to the hypothesis of the Real points up a difficulty which
is the central concern of my thesis to resolve, that is, whether the pragmatic maxim commits
Peirce to the Realism which he professed or to a form of Idealism. Peirce's response is
composed of more than two decade's work and my interpretation of it will concern his
writings​ ​on​ ​the​ ​normative​ ​sciences​ ​and​ ​the​ ​metaphysics​ ​of​ ​evolution.

As I have stated, Peirce initially introduced the pragmatic maxim as a methodological


tool in the process of inquiry. We ended our discussion of semiotic by recognizing the
distinction between those inferences which we can control and those we cannot. Unconscious
inferences constitute a large measure of the work of inquiry but as we saw, if inquiry is to
achieve its aims, if the fruitful development of a sign is to shortcut the interminable series of
interpretants which can represent it, it is to those inferences which we can control that
attention should be paid. Peirce regarded the traditional understanding of induction and
deduction to be more or less sufficient for the role they play in the control of inquiry. But to
recall, it was with abduction, or the formation of hypotheses, that Peirce identified the
fruitful, progressive aspect of inquiry. It is the deliberate control of this activity with which
Peirce is largely concerned, and thus the question formed by the end of the last chapter was --
among those hypotheses pertinent to the explanation of any given phenomenon, which are to
be​ ​admitted?

As we have seen thus far in this chapter the answer is defined in the imperative of the
pragmatic maxim. The maxim tells us nothing about which hypothesis is right, but aids only
in the economizing of inquiry by distinguishing content and thus making us aware that two
apparently different hypotheses may really be saying the same thing. To re-emphasize, Peirce
conceived of the maxim as merely methodological, a tool. Speaking of those, like James, who
misconstrued​ ​the​ ​intent​ ​of​ ​the​ ​maxim​ ​he​ ​says,

70
... one of the faults that I think they might find with me is that I make pragmatism
to be a mere maxim of logic instead of a sublime principle of speculative
philosophy​ ​(5.18).

In part he accepted responsibility for the misunderstanding saying that his initial articulation
of the maxim was too crude. The entire affair though is a good illustration of the application
of the maxim and the value of the community of inquirers. The meaning of the maxim is
explicable in terms of the conceivable practical consequences that belief in it would entail.
These consequences were manifest in the opinions of James and others and thus Peirce
became aware of the need to amend it so as to accord more closely with what he intended.
The original sign of the pragmatic maxim developed and grew as a consequence of the
interaction​ ​between​ ​a​ ​community​ ​of​ ​inquirers.

Peirce's refinement of it, though, is what presently concerns us. The "speculative
philosophizing" which the maxim engendered was the formation of a nominalist position
with regard to the purpose of thought. By emphasizing the practical consequences James
made concrete action (Secondness) to be the "be-all" and "end-all" of thought. Peirce,
however,​ ​demurred.

. . . I seem to myself to be the sole depositary at present of the completely


developed system, which all hangs together and cannot receive any proper
presentation in fragments. My own view in 1877 was crude. Even when I gave my
Cambridge lectures I had not really got to the bottom of it or seen the unity of the
whole​ ​thing​ ​(8.225).

Peirce recognized the need to broaden the context in which pragmatism is situated if it was to
avoid the nominalist interpretation of his initial formulation. The unity of which he speaks
above results from the grounding of the maxim in the context of the normative sciences. In
Peirce's analysis of inquiry pragmatism is the rule by which logic operates, but after years of
reflection​ ​he

found logic in and of itself to be insufficient for a full accounting of inquiry. Logic is a
normative guide in the practice of inquiry but, he claimed, is itself grounded in the more
fundamentally normative science of ethics. Ethics in its turn is ground in the science of
aesthetics,​ ​by​ ​which,​ ​he​ ​said,​ ​“I​ ​don't​ ​mean​ ​milk​ ​and​ ​water​ ​and​ ​sugar"​ ​(8.255).

71
By conceiving of pragmatism and logic as the fruitful term of more fundamental
inquiries in ethics and aesthetics he hoped to achieve a still higher grade of clearness of
thought. The third grade of clearness effected by employment of the maxim focuses too
specifically on action and, as Peirce saw, can mislead the proper attention of inquiry. As he
says,

The doctrine [the initial formulation of pragmatism] appears to assume that the end
of man is action -- a stoical axiom which, to the present writer at the age of sixty,
does not recommend itself so forcibly as it did at thirty. If it is admitted, on the
contrary, that action wants an end, and that the end must be something of a general
description, then the spirit of the maxim itself, which is that we must look to the
upshot of our concepts in order rightly to apprehend them, would direct us toward
something different from practical facts, namely, to general ideas, as the true
interpreters​ ​of​ ​our​ ​thought​ ​(5.3).

Mere action in itself Peirce characterizes as an instance of Secondness, a brute, individual


fact, without relation and thus impotent as a carrier of meaning. Action is a valid criterion of
thought but it is not its purpose. For Peirce the true upshot of thinking lies in the
establishment of rules of action, or as is by now familiar, habits of inference. A habit is a
regularity, a Third, and insofar as it regulates future events, is a general. This general is the
end which Peirce describes above as that which action wants, as that which invests action
with meaning. This still higher grade of clearness though is not explicable in terms of logic
alone.​ ​This​ ​requires​ ​investigation​ ​into​ ​the​ ​trio​ ​of​ ​what​ ​Peirce​ ​called​ ​the​ ​normative​ ​sciences.

72
Chapter​ ​Four

Normative​ ​Science

Though logic in itself is insufficient to the task of the fuller explication of meaning, it is
an element in the trio of Peirce's normative sciences. As the study of correct reasoning logic
prescribes rules for valid deductions and inferences, and as we have seen, the pragmatic
maxim aids logic in determining the admissibility of hypotheses. Peirce conceived of
normative science in general as the study of the "universal and necessary laws of the relation
of Phenomena to Ends" (5.121). The normative concern of logic then is with those
phenomena whose end is to represent something. It prescribes rules for valid inference, and
thus for true representation. As a guide then in the use of reason it is surely competent, but
reason itself is only a part of a much larger class, that of human action in general. As Peirce
said above, action wants an end, and the science which provides normative rules for the
relation​ ​of​ ​things​ ​whose​ ​end​ ​lies​ ​in​ ​action​ ​is​ ​what​ ​he​ ​called​ ​ethics.

Arriving at this conception of the role of ethics was difficult for Peirce because he
labored for years under the assumption that ethics was concerned simply with morality, with
right and wrong, and that it was more like an art, in the Aristotelian sense, than a science. He
came to see however that right and wrong are conceptions of ethics and, far from defining
ethics, they require the normative justification of ethics itself, as a science, in order to have
any​ ​meaning​ ​as​ ​categories​ ​of​ ​action​ ​at​ ​all.​ ​As​ ​Peirce​ ​says,

The fundamental problem of ethics is not, therefore, what is right, but, what I am
prepared deliberately to accept as the statement of what I want to do, what am I to
aim​ ​at,​ ​what​ ​am​ ​I​ ​after?​ ​.​ ​.​ ​.​ ​It​ ​is​ ​Ethics​ ​which​ ​defines​ ​that​ ​end​ ​(2.198).

This struggle is reflective in general of a fundamental tenor in Peirce's thought


characterized by a search for normative guides in the analysis of thought, or to put it in
Kantian terms, for the transcendental elements of cognitive activity. He eschewed the
conception of logic in the psychologistic terms of his day, as a discipline merely descriptive
of the way the mind works. As we have just seen he came to see ethics as needing a similar
reconceptualization, as being normative rather than descriptive, and his analysis of aesthetics
will bear the same fruit. esthetics is not about the beautiful and the ugly but about something
more​ ​fundamental​ ​than​ ​either.

73
At the root of his realization of these distinctions is, I think, his overcoming of the
persistent Cartesian misconceptions about the nature of consciousness which informs
traditional philosophical thought about logic, ethics, and aesthetics. Peirce says, "The
beautiful is conceived to be relative to human taste, right and wrong concern human conduct
alone, logic deals with human reasoning" (5.128). All three can be said to be the study of
human deliberation and judgment, and thus of mind, but the mistake is to continue thinking
of mind or thought as something residing in consciousness. As Potter has aptly summed up
this​ ​insight,

. . . thought is in consciousness rather than consciousness in thought. Mind is


thought, and thought is Thirdness, and Thirdness is ubiquitous. The human mind is
only one manifestation of Mind, the highest perhaps because it has the greatest
capacity for self-control, but not unique. Here again Peirce is insisting upon the
continuity of reality. If mind is anywhere, it is everywhere in one form or another.
41

This reminder of Peirce's conception of mind makes clearer, I believe, his discussion of
normative​ ​science.

To return to ethics, the dependence on it of logic becomes more apparent. If thinking is


to be deliberate and subject to control it must conform to some ideal, some end by which
comparison and judgment can be made. Furthermore, as a form of activity, thinking must
have an end. As the science which prescribes the ends of action, Peirce shows then that ethics
must precede logic in the normative structuring of inquiry. As he says, ... it is, therefore,
impossible to be thoroughly and rationally logical except upon an ethical basis" (2.193). Thus
ethics​ ​guides​ ​logic​ ​by​ ​providing​ ​the​ ​standards,​ ​the​ ​ends,​ ​to​ ​which​ ​logic​ ​directs​ ​thought.

But ethics itself relies on an even more fundamentally normative science, aesthetics.
Considering how Peirce perceives the categorial structures of these sciences will provide an
initial point of discussion for aesthetics. Logic is concerned with guiding the mediation
between sign and object in the representative function of the interpretant. Logic then, and
reasoning as such, is of the category of Thirdness. Ethics is of the nature of Secondness. It
connects two elements with one another, conforming action to an ideal. Aesthetics then is
characterized by the category of Firstness. It is conceived as such because its concern is with

41
​Vincent G. Potter, ​Peirce's Philosophical Perspectives​, ed. Colapietro, Vincent M., (New York: Fordham University
Press,​ ​1996),​ ​p.​ ​43.

74
ends as such, or more properly with the end, the ultimate standard by which all conduct can
be judged. This ultimate end is what Peirce characterizes as the summum bonum.
Consideration of this end qua end can be related to nothing else, but only as it is per se. As
Peirce​ ​says,

Ethics asks to what end all effort shall be directed. That question obviously
depends upon the question what it would be that, independently of the effort, we
should like to experience. But in order to state the question of aesthetics in its
purity, we should eliminate from it, not merely all consideration of effort, but all
consideration of action and reaction, including all consideration of our receiving
pleasure, everything in short, belonging to the opposition of the ego and the
non-ego​ ​(2.199).

Aesthetics prescribes to ethics the conditions under which an end can be an end as such, that
is, what makes an end desirable and admirable per se, without regard to any other
consideration.

What the normative sciences provide is a measure by which, as Peirce quaintly says,
"the sheeps may be separated from the goats," or that something can be judged either good or
bad. But such an act is always according to a standard. Logic says that a particular way of
thinking is good. It can say this because of the more general formulations of ethics. Ethics
says that a particular end is good and thus provides for the conformity of action to it. But it
can do this only because of the still more fundamental formulations of aesthetics. However
there must be some end to this legislative justification or else there would be no normativity
to these disciplines at all. Logic and ethics are ultimately based on aesthetics because the
regression​ ​stops​ ​here.​ ​Aesthetics​ ​seeks​ ​no​ ​further​ ​justification​ ​for​ ​the​ ​conditions​ ​it​ ​prescribes.

The question aesthetics concerns itself with is what it is possible to admire


unconditionally, without any further standard or ideal. More pointedly it "undertakes to
define precisely what it is that constitutes the admirableness of an ideal" (5.36). From such a
determination we can proceed to the ethical question of what, unconditionally, can be
adopted as an end of conduct, and then to the logical question of what can be adopted as an
ultimate end for the direction of inquiry. The role of aesthetics is vital to Peirce's purposes
here because ultimately it alone can avoid the conception of logic given by psychologism. It
is the lynch pin by which standards of inquiry can hold objectively not just for ome group of
people​ ​b​ ​u​ ​t​ ​for​ ​all​ ​rational​ ​agents.

75
The role of the aesthetic can be described in terms very similar to those Kant used in the
Critique of Judgment. First, as with ethics, Peirce saw that aesthetics should be defined in no
way by the terms beautiful and ugly. These rather are determinations of aesthetics, categories
which are defined in terms of the standards aesthetics prescribes. In fact, something may be
aesthetically good even though we may judge it to be ugly. This is so because of the way in
which​ ​Peirce​ ​formulates​ ​the​ ​criterion​ ​of​ ​aesthetic​ ​quality,

An object to be aesthetically good, must have a multitude of parts so related to one


another as to impart a positive simple immediate quality to their totality; and
whatever does this is, in so far, aesthetically good, no matter what the particular
quality​ ​of​ ​the​ ​total​ ​may​ ​be​ ​(5.132).

Thus, though we may be repulsed by a particular sight or deem it ugly, it is nonetheless


unified so as to impart an immediate quality of feeling. Whatever the latter may be it is
immaterial to the conditions which give of something an aesthetic quality. In this sense then it
makes no sense to speak of aesthetic badness or goodness at all, for either something imparts
a​ ​quality​ ​of​ ​feeling​ ​or​ ​it​ ​does​ ​not.

But the question of aesthetics is to define the conditions by which an ideal is admirable
in itself without further justification, thus giving conduct in general and inquiry in particular
the normative guides they need to achieve their purposes. The definition given above might
seem to serve that purpose but actually it does not. In fact the very idea of definition is
foreign to the work of aesthetic judgment as Peirce would understand it. To speak of
definition is to return to the abstract, rational work of logic. Definitions are relational
formulations entirely indicative of the mediating category of Thirdness. To recall, aesthetics
corresponds to Firstness. Its judgments are not the result of rational thought but rather of
feeling. The definition given above then refers not to the conditions by which an ideal is
admirable in itself but to the feeling which is the object of aesthetic experience. Aesthetic
judgment defines the admirable not through abstract definition but, as we will see, by the
formation of habits of feeling. Just what Peirce means by this will be clarified by returning
briefly​ ​to​ ​Kant.

76
Peirce​ ​and​ ​Kant​ ​on​ ​Aesthetics

As Peirce conceived of the role of the aesthetic, so Kant, in his Critique of Judgment,
saw aesthetic judgment as providing a bridge between theoretical and practical reason,
between the worlds of necessity and freedom. In the Critique of Pure Reason the
understanding provides the a priori categories by which phenomenal reality is known. With
regard to the domain of desire it is the practical employment of pure reason which legislates
imperatives for action. Finally, that faculty of the mind concerned with feeling is legislated
by​ ​means​ ​of​ ​what​ ​Kant​ ​called​ ​judgment.

The activity of judgment is neither that of the understanding nor of reason but one
mediating between the two. Certainly the understanding issues judgments but they are of a
different nature than those concerned with aesthetic apprehensions of feeling. The distinction
Kant makes here is between determinant and reflective judgment. Determinant judgment is
that whereby a particular is brought under a universal. For instance if the universal is that all
men are mortal, then any particular man, such as Socrates, can be found to be a case of the
universal. Reflective judgment is the reverse. A particular is given, and it is the purpose of
judgment​ ​to​ ​find​ ​the​ ​universal."42

I have been unable to find reference in Peirce to his having read Kant's Critique of
Judgment but the distinction defining the role of aesthetic judgment could not correspond
more closely. For Peirce as well aesthetic judgment is concerned with universals, with those
ideals which can serve as ends for conduct. However its judgments are formed not
determinantly but reflectively, and not in one abstractly-defining fell swoop, but over time
with all three of the sciences interacting. Aesthetic judgments begin with a particular,
whether it be a painting, an emotion, or more pertinently, a habit of inference, and help to
define ideals of conduct through what Peirce called the formation of habits of feeling. But
exactly​ ​how​ ​does​ ​this​ ​happen?

In any concrete situation thought progresses through the mediation of habits of


inference. Some of these habits describe well sedimented patterns of inquiry and are
generally stable, while others are not. In situations where experience compels us to review or
criticize any particular habit of inference we make changes appropriate to modify future
behavior. It is in this adaptive process of modification that aesthetics plays its normative role.

42
​ ​Reflective​ ​judgment​ ​can​ ​be​ ​seen​ ​here​ ​to​ ​model​ ​Peirce's​ ​notion​ ​of​ ​abductive​ ​inference.

77
It does this based on our capacity or experience aesthetically any action or object. A belief or
a habit of inference is a characteristic way of doing something that brings about in its doing a
quality of feeling, a liking or an attraction, with regard to it. When at some future time we act
on this habit, and thus conform to that particular way of acting, we will recall it having been
accompanied by a certain quality of feeling. It is the recognition of the pleasure attendant on
this feeling that acts as the rudder guiding thought toward ideal ends. Comparison is made
between a current situation and an earlier resolution such that their conformation results in a
quality​ ​of​ ​feeling​ ​that​ ​we​ ​find​ ​pleasurable.

But for Peirce this describes no mere hedonism. He says for "conduct to be thoroughly
deliberate, the ideal must be a habit of feeling that has grown up under the influence of
self-criticism" (1.574). Aesthetics becomes a normative guide only when it is brought into
interaction with the critical reflection of logic. As I have said, if the sphere of aesthetics were
considered apart from the influence of logic and ethics, it wouldn't be normative at all. It
would simply present experience with a series of aesthetic qualities. It is upon just such a
basis that hedonism is defined. Without critical reflection aesthetic experiences become
merely matters of taste, to use Kant's terminology. For the hedonist, whatever is found to be
pleasing becomes the basis for an ideal of conduct. Kant actually termed this the "taste of
sense" and distinguished it from what he called the "taste of reflection." For Kant it is by the
latter​ ​that​ ​one​ ​can​ ​form​ ​aesthetic​ ​judgments​ ​which​ ​can​ ​make​ ​claims​ ​to​ ​universality.

There is a subtle point of difference though between Kant and Peirce concerning the
role of aesthetic judgment and the claims to universality that it makes. Does Kant mean
simply to give an account of how we can judge something to be beautiful? Or do aesthetic
judgments have deeper, metaphysical implications? Though Peirce was critical of Kant's
work in the first Critique, the similarity in the overall architectonic of both thinkers is
striking. What Peirce discerned as the untenable conceptual structure of the first Critique is at
least implicitly amended in Kant's work in the second and third Critiques and parallels in no
small measure what Peirce seems to be attempting with the normative sciences. That is to
say, I see the relations between logic, ethics, and aesthetics as conceptually unified in much
the same way as the Critiques of Pure Reason, of Practical Reason, and of Judgment. In
addressing the question of the nature and real meaning of aesthetic judgment for Kant, and
the manner in which Peirce differs, I want to follow out this train of similarity so as to better

78
clarify​ ​their​ ​difference.

Kant's three Critiques are unified by a concern he terms transcendental, that is, they set
out to elucidate the conditions by which the operation of our various faculties are possible.
Pertinent to this elucidation is the determination of those elements of the operation of a
faculty which are a priori. The concern of the first Critique was how theoretical knowledge of
nature is possible. He answered this by describing how the categories of thought' give laws a
priori to phenomenal reality. The second Critique concerns the governance of desire, and
there Kant laid out the a priori principles of pure reason (in its practical employment) which
legislate with regard to desire. In the introduction to the Critique of Judgment Kant refers to
the gulf separating these two fields of human concern. On the one hand is the employment of
pure reason, the operation of which is constitutive of our knowledge of the world. This
defines the realm of necessity. On the other hand is practical reason, whose use is regulative
in the determination of desire. Its concern is with the field of human action and thus with the
realm of freedom. These two faculties of mind then concern sensible reality and supersensible
reality, Nature and Freedom, but between them there must be some principle of accord or
interaction which makes it possible for one to be affected by the other. This principle of
coherence​ ​Kant​ ​sought​ ​in​ ​the​ ​faculty​ ​of​ ​judgment.

As are the first two Critiques, the third is concerned with determining whether the
power of judgment has its own a priori principles and if so what they determine and the
manner in which they do so (constitutively, regulatively or otherwise). Kant finds that there is
an a priori element in judgment, however the kind of judgment he's concerned with here, as I
have mentioned, is reflective, and not determinant. It is not subsumptive but moves rather
from the particular to the universal and as such is not subject to the constitutive limits of the
categories of the understanding. Its work is evidenced in the course of scientific inquiry when
general laws are sought which explain a particular range of phenomena. The a priori element
Kant discerns is the presupposition in this activity which regards nature as an intelligible
unity. As Kant says, "the special empirical laws . . . must be considered . . . as if an
understanding which is not ours had given them for our powers of cognition, to make
possible a system of experience according to special laws of nature"43. This assumption of the
ultimate intelligibility of nature is not itself derived empirically and so in this sense it is a

43
​ ​Immanuel​ ​Kant,​ ​Critique​ ​of​ ​Judgement​,​ ​xxvii;​ ​Bd.,​ ​p.​ ​18.

79
priori. However it is different from the a priori principles of the understanding. The latter are
necessary conditions for the possibility of any and all objects of experience. The a priori
element of reflective judgment is rather " . . . a necessary heuristic principle which guides us
in our study of the objects of experience."44 It says nothing dogmatic about Nature itself but
rather​ ​that​ ​any​ ​investigation​ ​into​ ​nature​ ​involves​ ​this​ ​assumption​ ​as​ ​a​ ​principle​ ​of​ ​inquiry.

When one comes thus to make an aesthetic judgment the unified purposiveness of
nature is represented in our apprehension of the object. The basis of this representation is the
subject matter of reflective judgment. For pure reason the subject matter is phenomenal
reality, for practical reason, desire. For judgment what is legislated for is feeling. In judging
an object reflectively there is an accordance between its form and our cognitive faculty such
that the representation is accompanied by a feeling which is pleasurable. An object so judged
Kant calls beautiful. This judgment though regards nothing adhering in the object as such but
is rather "the feeling (of the inner sense) of that harmony in the play of the mental powers, so
far as it can be experienced in feeling.45 Based as it is in feeling, aesthetic judgment "is one
whose​ ​determining​ ​ground​ ​cannot​ ​be​ ​other​ ​than​ ​subjective."46

Yet Kant makes the claim that in making such judgments we speak with a universal
voice. It seems a paradox to claim that the exercise of a faculty is entirely subjective yet
universally applicable. In order to justify this claim to universality Kant has to demonstrate
some indispensable quality of aesthetic judgments without which our ability to judge,
reflectively or determinantly, would be impossible. He finds the basis for this justification in
the fact of our ability to communicate our judgments to others, " . . . the subjective conditions
of this faculty of aesthetic judgment are identical with all men in what concerns the relation
of the cognitive faculties . . . This must be true, as otherwise men would be incapable of
communicating their representations or even their knowledge.47 Communicability is thus a
condition of cognition and knowledge in general, and its fact in our experience is what allows
Kant to regard the subjective conditions of judgment to be the same for all. If this is true for
cognition in its employment by pure reason, it is equally true of the faculty of reflective
judgment, for it employs the very faculties used in determinant judgment, that is, the

44
​ ​Frederick​ ​Copleston,​ ​S.J.,​ ​A​ ​History​ ​of​ ​Philosophy.​ ​(New​ ​York:​ ​Doubleday.​ ​1960),​ ​vol.​ ​6.,​ ​p.​ ​352.
45
​K​ ant,​ ​60-1;​ ​Bd,​ ​p.​ ​90.
46
​ va Schaper, "Taste, sublimity, and genius: The Aesthetics of nature and art" In ​The Cambridge Companion to Kant​, ed.
E
Guyer,​ ​Paul,​ ​(Cambridge:​ ​Cambridge​ ​University​ ​Press,​ ​1992),​ ​p.​ ​375.
47
​ ​Ibid.​ ​p.​ ​377.

80
understanding and the imagination. In the aesthetic contemplation of an object these two
faculties are disengaged from their respective employments and achieve in their interaction or
"play" a harmony upon which is attendant a felt satisfaction of pleasure. It is the condition for
this harmony which is the same for all people and which Kant believes gives judgments of
taste​ ​the​ ​claim​ ​to​ ​speak​ ​with​ ​a​ ​universal​ ​voice.

But is there anything further to this ability of ours to judge something beautiful?
Beyond illuminating questions about cognition and communicability, is there something
more to aesthetic judgment than merely an accounting of taste, interesting as that may be?
There is. As I have mentioned, in his introduction Kant sought to provide a bridge between
the realms of Necessity and Freedom by investigating the role of aesthetic judgment. The first
Critique left our judgments about nature secure, however with the proviso that it is only
phenomenal reality that we know. The noumenal realm of things-in-themselves is left
undetermined by pure reason. For Kant though, aesthetic judgment provides access to it.
Though it neither constitutes objects of knowledge, nor legislates for desire in the realm of
action, aesthetic judgments represent nature as a phenomenal expression of the noumenal
realm. This is the deep meaning of beauty for Kant, that works of art, natural or otherwise,
reveal the purposive unity of nature which seems sundered by the apparent gulf between
reason and desire. In aesthetic judgment noumenal reality becomes determinable by human
faculty, and thus the faculty of judgment makes possible the transition from the domain of the
concept​ ​of​ ​Nature​ ​to​ ​that​ ​of​ ​the​ ​concept​ ​of​ ​freedom.48

As I mentioned, there is much similarity to be found with Kant in Peirce's account of


the normative sciences. The three Critiques and the three normative sciences have
respectively corresponding concerns -- thought, action, and feeling. However the context of
their expression differs because of Peirce's epistemological transformation of Kant. Peirce's
critique of meaning relocated the terminus of thought from a discrete, wholly sufficient act to
the convergence of opinion in the infinite long run. For Kant, the Real, at least in its
phenomenal expression, is always completely available to cognition whereas for Peirce it is
the term of an infinitely continuous process of inquiry. The Real then is not conceivable as
something which can be ultimately known in a completely determinate sense. Thus, as irony
would have it, Kant's things-in-themselves resurface in Peirce in the form of, as Apel has

48
​ ​Kant,​ ​Critique​ ​of​ ​Judgement​,​ ​(Cambridge:​ ​Hackett,​ ​1987),​ ​p.​ ​18.

81
commented,​ ​"paradoxes​ ​of​ ​the​ ​infinite."49

However, as I will show when discussing Peirce's metaphysics in connection with chaos
theory, this paradox is not a handicap in the sense that it was for Kant. For Kant, aesthetic
judgment provides some measure of overcoming the difficulties of the First Critique by
providing access to the noumenal realm. In glimpsing the purposive unity of Nature practical
reason is able to avail itself of the regulative hopes necessary to achieve its ends in the world
of Necessity. For Peirce, though, there is no noumenal realm and thus the role of aesthetic
judgment is somewhat different. As I mentioned, the noumenal realm for Peirce would be the
convergence of opinion in the infinite long run and it is the activity of thought guided by
logic which brings this about. But just as pure reason by itself leaves the noumenal realm
undetermined, so logic, without some normative guide, remains incapable of moving thought
forward.

The difference to be found between Kant and Peirce is that for Peirce aesthetic
judgment is an integral element in cognition in general whereas for Kant it is completely
unrelated to the activity of pure reason. For both thinkers it serves to give wider scope to the
limits of logic as such but Peirce makes of aesthetic judgment a moment in the process of
cognition itself. As we saw earlier Peirce regarded abduction as the only logical function
which introduces anything new into thought and by which it expands and grows. Like Kant's
reflective judgment, it moves from a particular and seeks a general by which it can be
explained.​ ​It​ ​is​ ​in​ ​this​ ​process​ ​that​ ​aesthetic​ ​judgment​ ​aids​ ​logic.

To reiterate, acting on a particular habit of inference, that is, engaging in any particular
way of thinking, has attendant upon it a quality of feeling that the doing of it brings about.
What aesthetics offers logical thought are qualities of feeling, nothing more. Ethics then,
consequent upon this feeling, regards it as admirable and thus defines it as an aim. It is in
reference to this aim that future conduct is judged. Upon reflection it will be seen that the
quality of feeling attendant upon the conduct conforms to the aim defined by ethics and will
thus be judged good. But the question of hedonism seems to come up again. Aesthetics offers
various qualities of feeling by which conduct can clearly be judged, but how can an ultimate
aim​ ​be​ ​determined?

49
​ ​Apel,​ ​p.​ ​146.

82
The​ ​Aesthetic​ ​ideal​ ​in​ ​Critical​ ​Deliberation

Peirce says, "If conduct is to be thoroughly deliberate, the ideal must be a habit of
feeling which has grown up under the influence of a course of self-criticism" (1.574). What is
this ideal feeling? It is not something that can be asked of aesthetics, for to do so would be to
introduce a standard of comparison, which is to move beyond the aesthetic and into the
ethical and logical. To recall, the aesthetic is of the category of Firstness, characterized by
immediacy and unrelatedness. Nothing more can be asked of aesthetics than to provide the
feeling​ ​of​ ​thought​ ​by​ ​which​ ​logic​ ​is​ ​able​ ​to​ ​judge​ ​its​ ​activity.

Whatever this ideal is it must be something ultimate and unitary, answerable to the long
term goals of inquiry. It must not be mutable under differing circumstances and it must be an
end unto itself, satisfied by no further motive beyond it. For Peirce this ideal can be nothing
other than Reason itself, or more precisely the development of Reason. Insofar as the
discussion concerns human cognition then he means by this human reason, but, as we saw
earlier, Peirce conceives of thought or reason as something not so much being in the mind as
the other way around, as mind being in thought. The development of Reason then is
equivalent to the development of Thirdness in the universe, to the formation and increasing
stabilization of habits of inference which bring about the furtherance of what Peirce called
"concrete reasonableness." Reason is the ideal which satisfies the above conditions because it
is always satisfied with itself. This is so because it is never completely embodied and
determinate. As he says, "The very being of the General, of Reason, consists in its governing
individual events. So, then, the essence of Reason is such that its being never can have been
completely perfected. It must always be in a state of incipiency, of growth . . . " (l.l65), and,
"The​ ​one​ ​thing​ ​whose​ ​admirableness​ ​is​ ​not​ ​due​ ​to​ ​an​ ​ulterior​ ​reason​ ​is​ ​Reason​ ​itself"​ ​(1.165).

Understood as such this ideal creates an end for conduct that Peirce characterizes as
"executing our little function in the operation of the creation by giving a hand toward
rendering the world more reasonable whenever . . . it is 'up to us' to do so" (l. 165). This is far
removed from a hedonist interpretation of Peirce's aesthetics and in fact reflects his
conviction that logic is rooted in the social principle. Abstractly, the ultimate ideal of conduct
has been identified with the development of reason, but more concretely Peirce said that the
ideal is a habit of feeling (which has grown up under the influence of a course of

83
self-criticism). The question then is what sort of habit of feeling is called for by this ideal
such​ ​that​ ​logic​ ​is​ ​able​ ​to​ ​"execute​ ​its​ ​little​ ​function?"

The short answer is love. For Peirce, a brief consideration of the nature of probability is
what brings him to this conclusion. All inferences made in the course of thought are subject
to ratios of probabilities. As he says, " . . . a given mode of inference sometimes proves
successful and sometimes not, and that in a ratio ultimately fixed" (2.650). He is not talking
about individual cases here where the notion of probability would have no meaning, for any
single inference is either true or false. It is only the mean result of a series of inferences by
which its validity as a mode of inference becomes known to us. Accordingly Peirce states
that, "there can be no sense in reasoning in an isolated case, at all" (2.652). The probability
inherent in a given mode of inference will manifest itself only after a great many risks are
taken using it and only at length will the result become clear. But humans are finite creatures
and the number of inferences we can make in a lifetime is limited. Thus we "cannot be
absolutely certain that the mean result will accord with the probabilities at all" (2.563).
Because of the way probability defines the limits of inferential thought the number of
inferences is indefinitely great and so relying on or thinking solely of one's self in one's
reasoning​ ​is,​ ​in​ ​the​ ​short​ ​time​ ​allotted​ ​us,​ ​doomed​ ​to​ ​failure.​ ​Thus​ ​does​ ​Peirce​ ​say,

Logicality inexorably requires that our interests shall not be limited. They must not
stop at our own fate, but must embrace the whole community. . . . This
community, again, must not be limited, but must extend to all races of beings with
whom we can come into immediate or mediate intellectual relation. It must reach,
however vaguely, beyond this geological epoch, beyond all bounds. He who would
not sacrifice his own soul to save the world, is, as it seems to me, illogical in all
his​ ​inferences,​ ​collectively.​ ​Logic​ ​is​ ​rooted​ ​in​ ​the​ ​social​ ​principle​ ​(2.654).

Peirce conceives the identification of one's own interests with those of an unlimited
community as a necessary prerequisite of logical thought. This sensibility Peirce sees
mirrored in the great religions of the world, especially as conceived by St. Paul in his
conviction of Charity, Faith, and Hope as the greatest spiritual gifts. That is, "interest in an
indefinite community, recognition of the possibility of this interest being made supreme, and
hope in the unlimited continuance of intellectual activity." Thus, odd as it sounds, reason
requires love in order for thought to be thoroughly logical. The ideal, which Peirce spoke of
as a habit of feeling, is in fact the feeling of love. Only can such a sentiment find self-identity
beyond​ ​the​ ​bounds​ ​of​ ​the​ ​ego.

84
The​ ​"New"​ ​Pragmatic​ ​Maxim

We have come a long way from the instrumental pragmatism of Peirce's initial
conception and the vulgar characterizations of it in the nominalism he despised. His attempt
to recuperate the maxim in the fuller context he intended has been the focus of my analysis of
the normative sciences. To what extent has he been successful thus far? Much as Peirce
criticized Kant, many believe his work in the normative sciences and metaphysics to be a
patchwork attempt to salvage the unsalvagable, that in trying to go beyond the instrumental
limits of the pragmatic maxim he engaged in the very kind of metaphysical speculation that it
was the point of the maxim to disabuse us of in the first place. His own summing up is given
in​ ​the​ ​following,

Pragmaticism makes thinking to consist in the living inferential metaboly of


symbols whose purport lies in conditional general resolutions to act. As for the
ultimate purpose of thought, which must be the purpose of everything, it is beyond
human comprehension; but according to the stage of approach which my thought
has made to it . . . it is by the indefinite replication of self-control upon self-control
that the ​vir50 is begotten, and by action, through thought, he grows an aesthetic
ideal, not for the behoof of his own poor noddle merely, but as the share which
God​ ​permits​ ​him​ ​to​ ​have​ ​in​ ​the​ ​work​ ​of​ ​creation​ ​(5.402).

Peirce reiterates here the original content of his definition of the maxim in which the meaning
of a concept is given in terms of how accepting its formulation as true will affect future
conduct. He in no way abandons the utility of the maxim as a means for the clarification of
ideas and as a practical aid in general for the work of inquiry. But then he continues and says,
"As for the ultimate purpose of thought . . . " What is this ultimate purpose? It is beyond our
comprehension and certainly something beyond the ken of the pragmatic maxim narrowly
defined to address. As he wrote to James in 1904, "I also want to say that after all pragmatism
solves no real problem. It only shows that supposed problems are not real problems. . . . The
effect of pragmatism here is simply to open our minds to receiving any evidence, not to
furnish​ ​new​ ​evidence"​ ​(8.259,​ ​emphasis​ ​mine).

The pragmatic maxim opens our minds and makes us receptive. As a tool of inference it
aids in the discrimination of possible interpretation and helps make of abduction the
uncannily efficient mode of inference that it is. But in terms of the end of thought, it furnishes
us with nothing. As Peirce said in the quote above, "it is by the indefinite replication of

50
​ ​Latin​ ​for​ ​“man”

85
self-control upon self-control that the vir is begotten." But self-controlled thinking is only
possible if there is some standard against which it can be judged, if some normative basis
exists by which thought can be guided from the outset. It is with a view to such a standard
that Peirce conceived the normative dimension of inquiry as embodied in the sciences of
logic,​ ​ethics,​ ​and​ ​aesthetics.

This is certainly not to say that deliberate self-controlled thinking has been going on for
millennia without any consideration of ends mediating its practice. What Peirce makes
explicit here is first, that thinking cannot be thoroughly logical without a consideration of
ends. As he says, "That which renders logic and ethics peculiarly normative is that nothing
can be either logically true or morally good without a purpose to be so. For a proposition, and
especially the conclusion of an argument, which is only accidentally true is not logical"
(1.575). The normative sciences make explicit what was only implicitly or vaguely assumed
before. Secondly, we can look at Mao Tse-Tung or Stalin or Hitler and find instances of a
manner of thinking guided by a very particular conception of ends. What Peirce's analysis of
normative inquiry elicits is the kind of end inquirers must adopt if thought is to fully achieve
its aims. Such an ideal, self-satisfied and wholly admirable in itself, is what Peirce
characterized as the development of Reason in its widest sense, that is, of Thirdness, or of
what he phrased "concrete reasonableness." Without a consideration of ends in this sense,
thought becomes a mere vehicle for the uncritical and idiosyncratic desires of pedagogues
and​ ​despots.

"The development of Reason," however, is an abstract formulation of this ideal. In order


for it to be efficacious for the practical work of inquiry it must be given in a conceptually
concrete framework. Inquiry investigates the Real, which, as we have seen, Peirce ultimately
identifies as the opinion of the community of inquirers in the long run. If this ideal opinion,
the Real, is the terminus of thought, then it must fully exhibit the characters of the ideal end
that Peirce characterized above as concrete reasonableness. But what is this? In short, it may
asked "What is the nature of the Real we are investigating?" Any specific investigation is
carried out within one of the many of the special sciences, and the norms which guide the
logic of this inquiry are described in general by Peirce's recuperated doctrine of
pragmaticism. The connection he saw uniting the latter with the former and answering the
question​ ​about​ ​what​ ​the​ ​Real​ ​is,​ ​is​ ​metaphysics.

86
A​ ​Prolegomena​ ​for​ ​Metaphysics

In places Peirce seems exceedingly sanguine about the work of metaphysics, thinking
for example that anyone with a passion for truth and a free afternoon ought to be able to clear
up most of the outstanding metaphysical issues.51 It is just such an attitude which seems to
have contributed to the widely divergent and often egregious flights of speculation which
Peirce began his thought by repudiating. It is in fact at this point in his philosophy that many
have shaken their heads, having judged Peirce to depart from and contradict his seminal work
in logic.52 But as Derrida has said, we cannot do away with metaphysics. It is as necessary for
thought as the air we breathe. Peirce would agree, however for him not just any metaphysics
will do. He described metaphysics as being in a "deplorably backward condition," a fact he
ascribed to its having been carried out by theologians and by uncritically nominalistic
scientists. "Find a scientific man," he says, "who proposes to get along without any
metaphysics . . . and you have found one whose doctrines are thoroughly vitiated by the crude
and​ ​uncriticized​ ​metaphysics​ ​with​ ​which​ ​they​ ​are​ ​packed"​ ​(1.129).

A metaphysics is necessary because it tells us about the structure of reality, knowledge


of which aids inquiry as it forms hypotheses about particular aspects of that reality. In the
process of forming and accepting hypotheses, especially as the concepts of science become
more general and abstract, the basic instincts which have guided abductive inference in the
quest for survival and the basic necessities no longer serve their purpose as adequately. A
metaphysics is a guide to theory formation and thus if it is not well considered our attempts to
achieve inquiry's aim will continue inefficiently, scuttling down every semiotic branch to its
logical​ ​end.

As Hookway has noted, "Peirce's synechism, the claim that we should favour
hypotheses that involve real continuity, exemplifies this sort of use of the metaphysics."53 At
the very least then Peirce is not so debonair about the importance of metaphysics. One of the
lessons he retained from his study of Kant is that metaphysics should be grounded on a sound
logical basis, and so it is from the fruit of all his investigation thus far in the logic of inquiry
and the normative sciences that a metaphysics adequate to the regulative needs of inquiry is

51
​ ​As​ ​quoted​ ​in​ ​Hookway,​ ​p.​ ​263
52
​W​ iener,​ ​1949.​ ​pp.84-85;​ ​Gallie,​ ​1952,​ ​ch.9;​ ​Almeder,​ ​1980,​ ​ch.​ ​IV
53
​ ​Hookway,​ ​p.​ ​264.

87
to​ ​be​ ​conceived.

Peirce recognized the highly speculative nature of metaphysics and would be the first to
remind us that his doctrine of Fallibilism applies equally and perhaps especially here.
However it is an investigation undertaken using the scientific method and not from the soft
cushion​ ​of​ ​the​ ​armchair.​ ​He​ ​says,

Metaphysics, even bad metaphysics, really rests on observations, whether


consciously or not; and the only reason this is not universally recognized is that it
rests upon kinds of phenomena with which every man's experience is so saturated
that​ ​he​ ​usually​ ​pays​ ​no​ ​particular​ ​attention​ ​to​ ​them​ ​(6.2).

Most metaphysical speculations are bad because they come, as it were, out of thin air,
conjured to drape over any number of theological, political, or social concerns, and not as a
result of logical investigation into thought itself. This latter concern is what illuminates the
"kinds of phenomena with , which every man's experience is so saturated." According to his
analysis thus far then Peirce will describe a picture of reality that will demonstrate the
universal applicability of his three categories of Firstness, Secondness, and Thirdness. He
will show that there are real laws that govern events and that they are so constituted as to
make explicable the affinity of cognition with the Real that it knows. A decisive ingredient in
this picture will be formed as a consequence of his doctrine of the Categories, that is, that
reality​ ​is​ ​evolutionary​ ​and​ ​not​ ​static​ ​in​ ​nature.

However, in preparation for a closer look at his metaphysics I want to introduce the
insights of chaos theory. As I mentioned in the introduction I conceive of my discussion of
chaos theory in two senses. First, it provides a scientific model of reality which concretely
illustrates the abstract characteristics of Peirce's metaphysics. Specifically, the concept of the
strange attractor in chaos theory provides a model for understanding how Peirce conceives of
reality as something causally efficient for the determination of hypothetical inference yet as
unbounded by a Newtonian determinacy and thus adequate to the infinite development of
Reason​ ​as​ ​characterized​ ​by​ ​the​ ​aesthetic​ ​ideal.

Secondly, I conceive of the following discussion as providing a limited and, let it be


said, fallible confirmation of Peirce's investigation of metaphysics. Now, the standard
objection to be made at this point, one that Peirce certainly would have made, is that no
philosophical theory can or should be grounded by the empirical results of any of the special

88
sciences. I could not agree more. For instance, the insights of quantum physics should not be
a mandate for the revision of philosophical theory. However a number of points should be
made. The first is that Peirce's metaphysics is in no way founded upon nor derives its validity
from the results of any of the special sciences but rather is conceived in terms mandated by
his categorial analysis and by the regulative ideals which make inquiry possible. It is rooted
in logic, not in empirical data. Secondly, in the scheme of his architectonic, metaphysics
comes last. It is the bloom of the flower, if you will, upon the bush and does not, logically,
play the more foundational roles that phenomenology and the normative sciences do. It is the
heuristic framework upon which inquiry evaluates hypotheses. Though it aids inquiry it is not
constitutive for it. And, unlike the normative logic of inquiry, metaphysics does not offer the
pragmatic​ ​maxim​ ​as​ ​a​ ​rule.​ ​Rather,​ ​as​ ​Apel​ ​has​ ​noted,​ ​it

takes as the object of inquiry the reality of the concepts postulated as habits by the
pragmatic maxim. Thus metaphysics becomes an investigation of the real process
of inquiry, which was for Peirce the conscious continuation of natural history and
so had to be, for him, a science; hence, it had to presuppose abductive hypotheses
which​ ​can​ ​be​ ​made​ ​subject​ ​to​ ​inductive​ ​testing​ ​by​ ​deducing​ ​their​ ​consequences.54

Apel continues by describing Peirce's conception of metaphysics as a "cosmological,


macro-empirical study whose vague, but heuristically indispensable, global hypotheses ​are
borne out or falsified by theory formation in the individual sciences.​"55 ​In fact, at 6.62, in
reference to his metaphysical theory of chance spontaneity, Peirce states his confidence that
his​ ​insights​ ​can​ ​be​ ​empirically​ ​confirmed.

Thus it is only natural, and in no way an instance of special pleading, that the results
of specific inquiry in the natural sciences be considered positively in our assessment of
metaphysics, at least in the limited sense in which they provide the reflective inquirer with a
confidence that the conception of the real guiding and constraining his or her inferences
manifests itself practically in the course of inquiry and thus is no mere wish or whim. To
reiterate what I said in the introduction, my discussion of chaos theory is given in the very
spirit of Peirce's epistemological program of inquiry -- the positing of an hypothesis, the
deducing of expected effects from the hypothesis, and the inductive testing of it in the course

54
​ ​Apel,​ ​p.​ ​144.
55
​ I​ bíd.

89
of our experience. If Peirce's metaphysical conjectures are true, then it is only reasonable that
evidence​ ​for​ ​it​ ​accrue​ ​in​ ​the​ ​course​ ​of​ ​scientific​ ​inquiry.​ ​To​ ​again​ ​quote​ ​Peirce,

Philosophy ought to imitate the successful sciences in its methods, so far as to


proceed only from tangible premises which can be subjected to careful scrutiny,
and to trust rather to the multitude and variety of its arguments than to the
conclusiveness of any one. Its reasoning should not form a chain which is no
stronger than its weakest link, but a cable whose fibers may be ever so slender,
provided​ ​they​ ​are​ ​sufficiently​ ​numerous​ ​and​ ​intimately​ ​connected​ ​(5.265).

I offer chaos theory not as a Cartesian-like indubitable basis for justification but as one
“fiber”​ ​aiding​ ​in​ ​the​ ​growth​ ​of​ ​knowledge.

90
Chapter​ ​Five

Chaos​ ​Theory

In 1890 Henri Poincaré published a paper in response to a challenge posed by the King
of Sweden. The king offered a prize to anyone who could find a solution to what was known
as the three-body problem. In classical Newtonian physics the orbit of one body around
another, for example that of the earth around the sun, can be precisely calculated. However,
consideration of a third body in the equation, in this case the moon, introduces complications
in the earth's gravitational field which makes precise measurement, according to Newtonian
formulas, impossible. Poincaré's solution was that there was no solution, at least not with
classical​ ​laws.​ ​He​ ​was​ ​given​ ​the​ ​prize.56

In a sense this was the beginning of what is known today as Chaos Theory. The
three-body problem is an example of the kind of phenomena chaos theorists are interested in
understanding, that is, simple, apparently deterministic, dynamical systems that display
unpredictable or chaotic behavior. What Poincaré showed was that no degree of fine tuning
the linear equations of Newton's laws can account for the behavior of a nonlinear system,
which is what the three-body problem is. In what follows I will define the various terms, such
as nonlinearity, that chaos theorists employ in their work. As for Poincaré, his insight
provided the initial theoretical conditions for understanding the behavior of nonlinear
systems, however it remained fallow for many decades, awaiting, through the discovery of
relativity and quantum theory, the advent of computers adequate to realizing new models of
analysis.

In its literal sense chaos theory is a bit of a misnomer, for although these scientists study
chaotic behavior, this behavior is not devoid of all sense of order. And though it is
characterized as a theory, as we will see it is not theory in the traditional sense. At its most
abstract chaos theory is about a new way of looking at the world. It does not offer any new
postulates or facts about the world, such as were ushered in by relativity theory and quantum
mechanics. It simply provides a way of accounting more fully than classical physics for the
behavior of dynamical systems. In discussing chaos theory I will show how it impels a
reconsideration of basic methodological and metaphysical assumptions behind traditional

56
​N. Katherine Hayles, Chaos Bound: ​Orderly Disorder in Contemporary Literature and Science, (Ithaca: Cornell
University​ ​Press,​ ​1990),​ ​pp.​ ​1-2.

91
science,​ ​particularly​ ​the​ ​broad​ ​conception​ ​of​ ​the​ ​universe​ ​as​ ​deterministic.

The kinds of things science studies are as perfuse and variegated as the universe itself:
planetary orbits, population fluctuations in field mice, heartbeats, water flow. In studying
these, science seeks an understanding of how they work, whether it be something as
individual as an atom or as complex as an ecosystem.What guides science in the search for
this understanding is the discovery of laws which govern how a thing or a system behaves.
Until very recently, the paradigm which governed the success of such understanding was
predictive ability. If we can predict the future state of a system, then in essence we
understand the laws which guide its behavior. For a wide range of phenomena science has
achieved this goal. Within certain parameters, often defined by human necessities and
practicality, science has given us a world ordered by predictability. Electricity lights our
homes,​ ​petroleum​ ​fuels​ ​our​ ​cars,​ ​and​ ​chemicals​ ​make​ ​us​ ​feel​ ​better​ ​after​ ​an​ ​indulgent​ ​meal.

An image that could characterize the way such a conception of science views the world
might be the clearing of a habitable place within a dense thicket. This image harks back
ultimately to the Cartesian split between mind and body and the scientific appropriation of it
during the Enlightenment as a mandate to conquer and tame wild nature to our purposes.
Within the clearing the world is ordered and predictable but at its edge that order begins to
break down. At the boundary of the clearing and beyond is the dense thicket, what traditional
science sees as waiting to be cleared away. This thicket in our image is the chaotic behavior
of systems which defy the traditional analysis of science. The way science has dealt with such
chaotic behavior is to regard it simply as complex information amenable to some future
technologically​ ​refined​ ​tool​ ​of​ ​analysis.

Chaos theorists would agree, but only to a point. Within this "clearing" science can
usefully predict a wide range of phenomena, yet disconcertingly, even the simplest systems
can begin to exhibit unpredictable behavior. This behavior is complex but does it mean that
scientists should look for more general and explanatory laws or that they should work on
refining their tools of analysis? Chaos theory answers no. As Poincaré seemed to suggest,
there is a limit to the accuracy with which the classical laws of physics can explain dynamical
systems. Why is this so? Let us turn to a closer look at just what chaos theory studies and
what​ ​it​ ​has​ ​to​ ​say​ ​about​ ​this​ ​complex​ ​information.

92
Definitions

Stephen Kellert has rather technically defined chaos theory as "the qualitative study of
unstable aperiodic behavior in deterministic nonlinear dynamic systems"57. Understanding
each term in this description will give a fairly coherent view of what chaos theory studies.
First, it is the study of dynamical systems. A system is any particular thing or process that a
scientist is interested in. It is composed of any number of variables which he or she identifies
and which thus define the parameters of the system. By assigning quantitative values to these
variables for a given time, the scientist can create a mathematical "picture" of the system. A
dynamical system is simply a mathematical model which describes the variation of this
picture over time. The variables comprising most dynamical systems change evenly and
continuously and thus are simply expressed using differential equations. Knowing the state at
one​ ​moment​ ​is​ ​sufficient​ ​for​ ​predicting​ ​the​ ​state​ ​at​ ​a​ ​future​ ​moment.

As mentioned in reference to the three-body problem, the kind of systems of interest to


chaos theory are nonlinear in nature. A linear system is one in which cause and effect are
related in a proportionate way. If one variable changes, then a corresponding and
proportionate effect arise at a future state in the system. In nonlinear systems there is no
simple relation between cause and effect. A change in one variable can disproportionately
affect the value of another such that for two variables with initially close trajectories, with the
onset of turbulence one may soar away in a manner unpredictable by classical physics. The
engine driving nonlinearity is what is known as iteration or the phenomenon of positive
feedback. The familiar chaotic squeal of a microphone brought too close to a speaker is an
example of iteration. As the system changes over time the variables feed back upon
themselves. Output becomes input and the repeated exponential multiplication of the
variables​ ​upon​ ​themselves​ ​sends​ ​the​ ​system​ ​into​ ​chaotic​ ​behavior.

Thus chaos theory is a qualitative study, for nonlinearity makes the tidy solutions
appropriate to linear systems impossible for nonlinear systems. Instead of taking a
quantitative approach to understanding the behavior of a system such that exact states of the
system in the future can be determined, chaos theory concerns itself with understanding
long-range​ ​behavior,​ ​with​ ​seeking​ ​patterns​ ​on​ ​a​ ​holistic​ ​rather​ ​than​ ​a​ ​reductive​ ​scale.

57
​ ​Stephen​ ​H.​ ​Kellert,​ ​In​ ​the​ ​Wake​ ​of​ ​Chaos​,​ ​(Chicago:​ ​The​ ​University​ ​of​ ​Chicago​ ​Press,​ ​1993),​ ​p.​ ​2.

93
Our definition is nearly complete. Though the behavior of nearly any dynamical system
can be described qualitatively, it is with regard to systems that are unstable and aperiodic that
chaos theory is concerned. An extremely simple example of a stable system would be a bowl
with a marble at the bottom. If you displace the marble and move it the edge of the bowl and
then let go, it will return to the bottom. It resists small perturbations in its equilibrium. An
unstable system on the other hand is one whose behavior does not resist small changes.
Rather than being absorbed, perturbations may lead to a future state where the system
exhibits chaotic behavior. Additionally, chaos theorists are concerned with aperiodicity. In
aperiodic systems variables never fall into a regular pattern of repetition but seem to wander
apparently randomly. Mathematically, the paragon of this is the mathematical value of pi: it
has no definite value and no repeatable pattern. So, behavior which is unstable and aperiodic
is, like the dense thicket, very complex. It has no repeatable pattern and manifests even small
changes​ ​to​ ​its​ ​equilibrium.

As Kellert described it then, chaos theory is the qualitative study of unstable aperiodic
behavior in deterministic nonlinear dynamical systems. The last term to discuss here is
deterministic. It seems oddly placed here among these others, but it is this very fact that
makes chaos theory a compelling field of inquiry. Chaos theory is not concerned with ཞི་
disparate, exotic branch of physical phenomena but rather with simple straightforward
dynamical systems such as water dripping from a faucet or the beating of a heart. These
processes can be described using very rigorous mathematically deterministic models. Yet
given certain conditions, such as a sudden downpour over a river, predictable behavior
suddenly becomes chaotic and unpredictable. Accounting for the rise of complex behavior in
simple​ ​ordered​ ​systems​ ​is,​ ​in​ ​the​ ​end,​ ​what​ ​chaos​ ​theory​ ​is​ ​concerned​ ​to​ ​provide.

Phase​ ​Space​ ​and​ ​the​ ​Mapping​ ​of​ ​Dynamical​ ​Systems

The manner in which chaos theory makes this accounting, as I have said, is qualitative.
Where traditional science crunches numbers into equations, chaos theorists draw a map. The
kind of map they draw is one corresponding to what scientists call a system's phase space.
Actually, assessing the behavior of a system by drawing a map of its phase space is a
technique common to a wide range of scientific disciplines. However it is the kind of space
chaos theorists have been able to conceive, win the aid of computers, that makes their

94
analysis different. The phase space of a system is an n-dimensional plotting of a sufficient
number of the variables comprising it so as to describe its movement, that is, how its
variables​ ​change​ ​over​ ​time.​ ​As​ ​Gleik​ ​says,

In Phase space the complete state of knowledge about a dynamical system at a


single instant in time collapses to a point. That point is the dynamical system -- at
that instant. At the next instant, though, the system will have changed, ever so
slightly, and so the point moves. The history of the system can be charted by the
moving​ ​point,​ ​tracing​ ​its​ ​orbit​ ​through​ ​the​ ​phase​ ​space​ ​with​ ​the​ ​passage​ ​of​ ​time.58

For instance, the trajectory of a rocket lifting off into outer space would have, as variables,
displacement and velocity. In real life the trajectory (the path of its flight) is a straight line,
but as mapped in phase space the trajectory twists and turns due to booster stages and the
varying effects of gravity. What phase space gives the scientist is a model for understanding
how the variables change over time. The plotted variables describe the "shape" of the
system's​ ​overall​ ​behavior.

The shape that researchers of dynamical systems look for is what they more technically
call an attractor. By defining the parameters of a system's attractor a scientist can predict
what future behavior of the system will be like. But what is an attractor? Like maps of phase
space, attractors are a normal part of the method of traditional scientific investigation. Before
the advent of chaos theory three general types had been identified in the study of dynamical
systems: fixed point, limit cycle, and torus attractors. A look at each in turn will prepare for
an​ ​understanding​ ​of​ ​the​ ​new​ ​kind​ ​of​ ​attractor​ ​which​ ​interests​ ​chaos​ ​theorists.

A ​fixed point attractor describes a system which is stable and rigorously periodic. An
example would be a pendulum swinging in a vacuum. If we define the variables as velocity
and​ ​displacement,​ ​then​ ​a​ ​plotting​ ​such​ ​as​ ​that​ ​in​ ​Figure​ ​A​ ​is​ ​what​ ​results:

​​​​​​​​​​​​​

58
​ ​James​ ​Gleick,​ ​Chaos:​ ​Making​ ​a​ ​New​ ​Science​,​ ​(New​ ​York:​ ​Viking​ ​Penguin,​ ​1987),​ ​p.​ ​134

95
Since friction is absent in this system the values for displacement and velocity remain
constant, that is, periodic. The behavior of this system is fixed; barring outside influence it
will never change.59 The circle inscribed by the variables on the graph is what scientists term
the attractor of this system. That is to say that its dynamics affect the variables such that their
values consistently converge on the phase space map of a perfect circle. If a shock were given
to the system and the displacement increased at some point, the attractor would still remain
unaffected; it would be larger but it would still be a circle. For a pendulum in the real world,
where friction plays a role, the attractor is also a fixed point, as in Figure B above. Starting
with an initial displacement and velocity the trajectory will spiral inward in phase space until
it meets the intersection of the two axes. This is the point toward which the system is
attracted,​ ​meaning​ ​that​ ​its​ ​behavior​ ​tends​ ​toward​ ​a​ ​fixed​ ​point,​ ​a​ ​-state​ ​of​ ​complete​ ​rest.

The next stage of complexity in system dynamics is defined by the limit cycle attractor.
Such a system tends not toward one state but cycles along a trajectory formed by two points.
A classic example of this is the predator-prey system found in wildlife populations. Take for
example the populations of trout and pike in a lake. If the populations of both start out
initially equal, over time the pike population will grow as they feed on the trout, and
correspondingly the trout population will decline. As the trout decrease the bloated pike
population will find their food supply limited and so some will begin to die off. As the pike
die off the trout slowly recover until the population of both species is again equal. Clearly,
the populations never achieve a fixed state but rather oscillate between population limits. For
either​ ​of​ ​the​ ​populations​ ​the​ ​attractor​ ​in​ ​phase​ ​space​ ​looks​ ​like​ ​a​ ​standard​ ​sine​ ​Wave.

Increasing the behavioral complexity even more results in a yet more sophisticated kind
of attractor. If we include in our frame of reference two limit cycles in interaction with each
other, the plotting of its dynamics in phase space produces an attractor with the mathematical
shape of a torus. In fact it is this kind of attractor which is used to model the gravitational
orbits of heavenly bodies such as planets. However, this brings us back to the introductory
remarks about Poincaré. For any two systems, such as two planets, in interaction with one

59
​In connection with this Alexander Argyros makes an interesting analogy with Derrida's criticism of metaphysics. "It is
tempting to speculate that what Derrida understands as metaphysics might simply be a system generating a normal attractor.
If by metaphysics is meant the desire to conceptualize the world as a system whose dynamics are reducible to a nontextual,
fully present essence, then it would indeed appear that stable attractors function as some sort of transcendental signified. Any
system whose behavior is guaranteed by an underlying determinable attractor would be onto theological insofar as the
attractor serves one of the traditional purposes of God - to control and channel the play of traces" (Alexander J. Argyros, ​A
Blessed​ ​Rage​ ​for​ ​Order​,​ ​(Ann​ ​Arbor:​ ​University​ ​of​ ​Michigan​ ​Press,​ ​1991),​ ​p.​ ​251).

96
another, the torus attractor is sufficient in the description of their behavior. But as Poincaré
demonstrated, further complexification, such as the introduction of a third body, skews the
results of traditional analysis and makes exact prediction impossible. In terms of phase space,
the kind of behavior manifested in the three-body problem cannot be described by the torus
attractor. This turbulence in the equations has traditionally been dealt with in a reductionist
manner, by breaking down the variables into an isolated calculus of coupled series and then
replotting them on the torus surface, hoping that adjustments in the equations would not
affect the overall stability of the attractor. To return to the metaphor of the clearing and the
thicket, this strategy has only a limited success. It makes predictions in the short run feasible,
but​ ​leaves​ ​those​ ​in​ ​the​ ​long​ ​run​ ​seemingly​ ​indeterminable.

The static Aristotelian perfection of the celestial realm has long since given way to the
conception of it as dynamic and changing. Poincaré's insight into the three-body problem
gave evidence of this, but even moreso it challenged the basic assumptions of the Newtonian
view of the universe as thoroughly ordered, deterministic, and predictable. Using the
clearing/thicket metaphor a detractor might persist in conceiving of turbulent chaotic
behavior as simply very complex information awaiting comprehension via refined analytical
tools. If this were true it would make what chaos theory has to say about the universe an
interesting topic of conversation, but hardly paradigmatic. II will address this issue shortly,
but first II want to discuss the kind of attractor chaos theorists have found to model chaotic
behavior. It may well be the very refined analytical tool the detractor hopes for, but it has
implications​ ​that​ ​will​ ​suggest​ ​a​ ​fundamental​ ​reconception​ ​of​ ​the​ ​dynamics​ ​of​ ​the​ ​universe.

Strange​ ​Attractors

As I have stated, what interests chaos theorists is understanding the dynamics of a


system that can go from ordered linearity, to turbulence and chaos. The paradigmatic
example is the flow of water in a river. Initially its flow can be completely deterministic, but
as its volume and velocity increase vortices and eddies appear, looping one within the other.
As the system is pushed further and further, all becomes a swirling turbulent confusion and
totally unpredictable. The increasing complexification can be modeled using the series of
attractors I've described above. Starting with a fixed-point attractor the flow jumps to a limit
cycle. From the limit cycle it transforms to trajectories describing the surface of a torus. From

97
here, if the Newtonian model were pursued, one would expect the torus to transform into
higher mathematical dimensions. Instead, what chaos theorists have found is that chaotic
behavior is modeled not by increasingly higher dimensions in phase space but rather by a
fractional​ ​dimension,​ ​that​ ​is,​ ​a​ ​space​ ​between​ ​two​ ​and​ ​three​ ​dimensions.

To illustrate this let me describe the pioneering work of Edward Lorenz, the father of
chaos theory. In 1960 Lorenz was using computers to help solve mathematical equations
modeling the earth's atmosphere. In doing a forecast he entered data for a variety of variables
and ended up with a prediction of the future state of the weather. At a later time, wanting to
clarify some details, he returned to that forecast and re-entered the data on the system's
variables. The first time, he entered numbers up to the sixth decimal place. This time however
he rounded off to just three. When he came back to check the results of the second run he
found​ ​a​ ​completely​ ​different​ ​forecast.​ ​As​ ​he​ ​concluded,

It implies that two states differing by imperceptible amounts may eventually


evolve into two considerably different states. If, then, there is any error whatever
in observing the present state -- and in any real system such errors seem inevitable
--​ ​an​ ​acceptable​ ​prediction​ ​...​ ​in​ ​the​ ​distant​ ​future​ ​may​ ​well​ ​be​ ​impossible.60

If the weather in the real world behaved like the computer model, weather prediction beyond
a​ ​few​ ​days​ ​would​ ​be​ ​impossible.

What Lorenz discovered is one of the defining characteristics of chaos theory, that
nonlinear dynamical systems display sensitive ​dependence on initial conditions. ​This concept
is illustrated in the ​now famous notion of the "butterfly effect", which is to say that the
flapping of a butterfly's wings in Argentina today could cause a tornado in Kansas tomorrow.
The image is perhaps a bit sensational but what it illustrates is that dynamical systems in
nature cannot be understood by isolating them from the dynamical system of the world as a
whole. In other words, conceiving of the world as the sum of its parts is no longer viable
because the parts are sensitively connected and dependent upon one another. The vision this
heralds​ ​is​ ​a​ ​holistic​ ​and​ ​dynamic​ ​one​ ​rather​ ​than​ ​a​ ​reductively​ ​deterministic​ ​one.

With this realization Lorenz began searching for another way to model the weather
system. Instead of a quantitative approach, the limits of which he had just witnessed, he tried

60
​ ​Edward​ ​Lorenz,​ ​"Deterministic​ ​Nonperiodic​ ​Flow"​ ​in​ ​Journal​ ​of​ ​the​ ​Atmospheric​ ​Sciences​ ​(20,​ ​1963),​ ​p.​ ​133.

98
a qualitative approach. Before this, meteorologists had used traditional equations which
produced multidimensional torus attractors, but the predictive ability it conferred extended
only to a few days. What Lorenz was able to do, with the aid of the enormous calculating
abilities of the computer, was to plot the complex trajectories of his nonlinear equations.
What resulted was one of the most fascinating discoveries ​of chaos theory - the strange
attractor:

The attractor is dubbed strange because it reconciles two seemingly contradictory


characteristics: it models behavior which is nonperiodic yet at the same time is bound within
a finite area of phase space. Again, nonperiodicity refers to the fact of a variable never
repeating itself in a patterned way. In phase space it means that the trajectory never crosses
itself but continues on infinitely. The strangeness adheres in its not being randomly strung out
in an infinite area of phase space but rather that the trajectories converge toward a definite
shape or basin of attraction. The dynamics here are akin to the taking of an infinitely long
string and somehow mushing it up into a finite space. How is this done? What kind of shape
can​ ​possibly​ ​satisfy​ ​such​ ​conditions?​ ​The​ ​answer​ ​lies​ ​in​ ​fractal​ ​geoтеtry.

The​ ​Fractal​ ​Dimension

The shape of a strange attractor is neither a fixed point, a sine wave, nor a torus. These
attractors are one and two-dimensional shapes that cannot satisfy our conditions. A
one-dimensional attractor obviously cannot; and on a two-dimensional surface it is possible

99
for trajectories to cross each other, thus making periodic behavior possible. But neither can
the attractor be three-dimensional. Any system in nature is dissipative, meaning that over
time it loses energy. As a system progresses this loss manifests itself in phase space as a
contraction in area. As Kellert has said, consequent upon this contraction "[t]he attractor
represents the shape that any initial set of points will approach asymptotically, so it must have
no volume in the three-dimensional state space. Therefore, the dimension of the attractor
must be less than three."61 But it must also be greater than two. The kind of shape a
nonintegral​ ​dimension​ ​describes​ ​is​ ​called​ ​a​ ​fractal.

The word 'fractal' comes from the Latin ​fractus​, which means ‘irregular’, and it was
coined by the mathematician Benoit Mandelbrot in an attempt to more adequately describe
the irregular geometry of the world he saw around him. A simple illustration of fractal
geometry is the jagged edge of a coastline. On a large map one could imagine taking a string,
laying it out among the various twists and curves, and then measuring the distance of the
string consumed using the scale at the bottom of the map. But this would be an inaccurate
measure, for if one moved in closer the straight lines that the map displays would exhibit
jagged detail too fine for the maps particular scale. At a closer range a second measure could
then be taken, but again, moving in to a closer scale would reveal more detail that we were
initially too far away to see. The point is that this maneuver can go on indefinitely. Where
before there was a smooth line, each magnification or change of scale reveals even finer
details. Perhaps the most interesting characteristic of fractal geometry is that each of its scales
is self-similar. The jagged edges of a pebble lying on the coast mirror the same kind of
jaggedness the coastline has when viewed on a map. The same is true of the branching of
blood​ ​vessels​ ​in​ ​the​ ​body,​ ​from​ ​the​ ​largest​ ​vessel​ ​down​ ​to​ ​the​ ​smallest​ ​capillaries.

This iterative nature of the fractal dimension is something Mandelbrot discovered when
he used a computer to iterate a basic algebraic expression, C^2 + Z. By starting with initial
values for Cand Z the computer is directed to reassign the result as the value of Cand then run
the equation again, ​ad infinitum​. Mathematically extrapolated, the result, modeled on a
computer, is the wildly fantastic spins and spirals which adorn the covers of many a book on
the subject of chaos theory. The situation is much akin to reflecting a mirror in front of
another mirror. The reflections, self-similar at increasingly smaller scales, seems to go off

61
​ ​Kellert,​ ​p.​ ​15.

100
into infinity. This is the way the strange attractor works. Within a fractal dimension it is able,
like​ ​a​ ​finely​ ​layered​ ​pastry​ ​puff,​ ​to​ ​weave​ ​infinite​ ​trajectories​ ​within​ ​a​ ​finite​ ​space.

Many different attractors with varying fractal dimensions have been found using this
approach of modeling nonlinear systems. To say that an attractor has a fractal dimension of a
particular value, say 2.6, is nothing more than to describe a geometrical object. Remember,
this geometrical object, the attractor, is a kind of map which qualitatively indicates how a
system's behavior changes over time. If we were to say that this map were two-dimensional,
and used a torus to illustrate it, we could easily see how the trajectories moving over this
familiar dimension describe the behavior of a particular system. If we were to say that this
map had a fractal dimension, it would be an indication that the shape of the attractor is
somewhere between two- and three-dimensional. Assigning a fractal value can be construed
as a way to measure an attractor's "intrusiveness" (as Kellert has described) into three
dimensional space. The strange attractor loops and folds these infinite trajectories into a finite
space and the fractal value tells the investigator how tightly it does this. The fractal value also
characterizes the scaling properties of the attractor, as well as indicating how the attractor
looks at increasing scales of magnitude. What we're looking at when we see those beautiful,
full-color, otherworldly agglomerations of spirals on dust jackets is a highly magnified view
of a strange attractor's structure. Such images have appeal firstly because they're beautiful to
look at. But this beauty derives in large part from the symmetry it exhibits. No matter how far
down one scales in magnification, the detail observed mirrors the detail of the structure at
larger magnifications. Contrary to popular and traditionally scientific conceptions of chaotic
turbulence as being random and without order, these fractal attractors exhibit a highly defined
hierarchy​ ​of​ ​order.

In a sense, chaos theory has cleared away a bit of the proverbial thicket. It has made
accessible to analysis what hitherto seemed incomprehensible. But the tool that has made this
possible, the computing ability of high speed computers, has revealed something other than
what was expected. Perhaps it is better to say that the thicket remains and that the access
gained is not so much dominatory as it is navigatory. Strange attractors furnish no equations
for the exact prediction of a system's future state but rather allow investigators an
understanding of how the system as a whole behaves. The approach here is holistic rather
than reductionistic and dispels the conception of chaotic behavior as anomalous white noise.

101
On the contrary, strange attractors show that there is method to the madness. Not only do they
assume a localized area in phase space, but analysis of their fractal dimensions reveals quite
orderly hierarchical self-similarity at all scales of its structure. It is this dimensional
characteristic which makes the conception of a strange attractor as a "bound infinity"
possible,​ ​and​ ​thus​ ​which​ ​makes​ ​the​ ​system​ ​it​ ​describes​ ​not​ ​so​ ​chaotic​ ​as​ ​once​ ​thought.

The phrase, "local unpredictability, but global stability" is often used to characterize
attractor analysis of chaotic systems. On the negative side it claims that we cannot make the
kind of predictions hoped for by traditional science. Taking the weather system as an
example, a system which is nonlinear and thus highly sensitive to initial conditions, we see
that highly accurate forecasts cannot be made more than a few days into the future. On the
positive side, however, plotting the weather system as a whole reveals globally predictable
behavior. It is a qualitative understanding we gain rather than quantitative. But of what
practical use is this kind of understanding, one may ask, if it can tell us nothing concrete
about the future. It is like a lens which allows us to see an object in the distance more clearly,
but it seems to do nothing about bringing us any closer. In fact, the insights of chaos theory
have provided scientists among a broad array of disciplines strategies for productively
channeling the dynamics of chaotic behavior.62 What is of more direct concern to my thesis,
however, is the way in which chaotic behavior, instead of being senselessly random, may be
the​ ​very​ ​engine​ ​of​ ​novelty​ ​and​ ​creativity​ ​in​ ​a​ ​dynamic​ ​and​ ​evolving​ ​universe.

Beyond​ ​Newtonian​ ​Determinism:​ ​Chaotic​ ​Systems​ ​and​ ​the​ ​Creative​ ​Vitality​ ​of​ ​the​ ​Cosmos

The sensitive dependence on initial conditions that nonlinear systems have may define
the limit of our predictive powers but it provides a way of accounting for the change and
growth we see in the world around us. For instance, on the biological level Darwin offered a
paradigm for the mechanism of variation, but was unable to offer an explanation for the
conditions - that allow for a gene to mutate. If the cellular dynamic were to be
reconceptualized on a chaotic attractor, it might be possible to conceive of the variation
which produces change and the holistic integrity of the system which maintains stability as
interactive and cooperative expressions of the underlying dynamic which chaos theory

62
​ ​See​ ​William​ ​L.​ ​Ditto​ ​and​ ​Louis​ ​M.​ ​Pecora,​ ​"Mastering​ ​Chaos",​ ​Scientific​ ​American​ ​(August​ ​1993)​ ​p.​ ​62-8.

102
studies.​ ​As​ ​Argyros​ ​has​ ​commented,

[Chaos theory] allows us to model, and thereby understand, a strategy that nature
has apparently chosen to deal with those of its problems most in need of creative
solutions. Chaos may be the strategy selected by evolution to accommodate the
dual demands of conservation and innovation, or, perhaps more accurately,
evolution may be the strategy chosen by a chaotic universe to wrest complexity
from​ ​the​ ​teeth​ ​of​ ​entropy.63
What these strange attractors seem to embody is a self-organizing principle in the world, and
imply that what we negatively perceive as chaos, instead of being random and without sense,
is harnessed as an agent of productivity and holistically stable change. It is much akin to how
the devil was portrayed in medieval history, "that nature does something against its own will
and,​ ​by​ ​self-entanglement,​ ​produces​ ​beauty."64

These thoughts are only speculative. My intention in discussing them is to suggest that
chaos theory offers a new model for understanding the universe, one which moves beyond
the Cartesian/Newtonian paradigm of traditional science, and one which, as I will show in my
final chapter, reflects with remarkable accuracy Peirce's metaphysical conception of the real.
So far in this discussion I have tried to demonstrate the novel and potent analytical strategies
developed by chaos theory, however at various points I have made assumptions about the
inadequacies of traditional methodologies which now need to be more fully addressed before
moving​ ​on.

The question concerns the image of the clearing and the thicket I introduced earlier, that
the slow paring away of the thicket into systems of predictable order is the best method we
have for understanding the world. Though chaos theory may provide an interesting way of
slicing up the picture, does that mean that the traditional view should be abandoned? To do so
would mean the reconceptualization of some fundamental metaphysical notions about the
nature of what science investigates, the warrant for which should not be considered without
forceful evidence. Closely bound to the clearing/thicket metaphor is the notion of
predictability and the deterministic assumptions which gird it. Limitations on this ability is
what is suggested by the work of chaos theorists. The critique of these assumptions then is
central​ ​to​ ​assessing​ ​the​ ​positive​ ​value​ ​of​ ​the​ ​insights​ ​of​ ​chaos​ ​theory.65

63
​ ​ ​Argyros,​ ​p.​ ​254-5.

64
​ ​Gleick,​ ​p.​ ​142.
65
​T​ he​ ​following​ ​discussion​ ​of​ ​determinism​ ​is​ ​largely​ ​indebted​ ​to​ ​Kellert's​ ​analysis.

103
As I have already made mention, chaos theory offers no new postulates about the
physical structure of the world, as did relativity theory and quantum mechanics. Einstein's
theory addresses the impossibility of collating events into an absolute temporal order and
quantum theory of assigning definite values in the measurement of particles. By describing
facts about the way the world is, these theories impose necessary limits on what we can hope
to understand. The question concerning chaos theory is predictability. Does the apparatus of
strange attractors, fractal dimensions, and nonlinear dynamics reveal limitations imposed by
facts of the world that chaos theory has discovered? No, it can make no such claim, yet its
ineluctable conclusion is that from classical, thoroughly Newtonian dynamical systems,
behavior​ ​can​ ​arise​ ​which​ ​eludes​ ​our​ ​best​ ​efforts​ ​at​ ​prediction.

The detractor claims that the improvement of predictive ability simply awaits advances
in technical analysis and would argue that abandoning the traditional paradigm because of
current limitations would be like abandoning Kepler's theory of planetary orbit because the
telescope hadn't been invented yet. The response of chaos theory is that, clearly, our ability to
predict future events is an eminently useful and necessary power, not only in the sphere of
practical human need but in the realm of scientific investigation as well. It is a measure by
which we can judge our explanation of physical phenomena to be accurate. But this accuracy
has its limits. Throughout the history of science, investigators have never pretended that their
equations predict with absolute precision. Scientists routinely round off the values of
variables in an experiment according to an acceptable range dictated by the constraints
involved. In large measure this procedure has not hindered confidence in the assumptions
guiding scientific inquiry, for science has made possible bridges that hold, airplanes that fly,
and chemicals that heal. Such feats can accept a degree of imprecision it the theories and
equations​ ​which​ ​make​ ​them​ ​possible​ ​and​ ​still​ ​viably​ ​address​ ​human​ ​need.

But there is some measure in which this procedure does hinder confidence in traditional
methodology. Whether it be the three-body problem, the weather system, or a fibrillating
heart, science has always encountered limits to its predictive ability. Now it is clear that given
a certain parameter of variables, imprecision in measurement has no appreciable effect in the
prediction of results. This is tacitly formulated in the methodological assumption that small
causes have small effects. However as Edward Lorenz discovered in modeling the weather
system, this is not true. Rounding off to six decimal places in the equation he was using,

104
rather than three, predicted completely different results. The explanation of this, as I have
stated, is the sensitive dependence on initial conditions that nonlinear systems exhibit. The
rounding-off involved in measurement introduces a degree of "vagueness" that grows
exponentially with the time frame involved. For a system such as the weather a small time
frame of say several hours is sufficient for a highly accurate prediction, but as the time frame
extends to a few days, a week, then a month, the variables involved grow exponentially and
become overwhelmingly complex. Making a useful, much less accurate, prediction involves
the​ ​decimal​ ​representation​ ​of​ ​values​ ​which​ ​for​ ​all​ ​intents​ ​and​ ​purposes​ ​are​ ​infinite.

The dogged detractor could still insist that, in principle, such a prediction is possible,
that a merely practical limitation should not be conflated with a theoretical one. But is the
limitation merely practical? The truth is that given the exponential growth of decimal
representation with increasing time frames, prediction tasks can theoretically be made so
difficult so as to require, literally, astronomical resources in order to calculate. Can we
predict where and with what intensity an earthquake will strike the United States in the year
2027? The consideration of every conceivable factor and variable involved would require for
storage more electrons than there are in the observable universe. By a ​reductio ad absurdum
the argument chaos theory makes transforms what seems to be a practical limitation into a
theoretical one, one based on our human finitude and one which thus has epistemic
consequences.​ ​As​ ​Kellert​ ​has​ ​stated,

Our finitude does not derive directly from any law or postulate of physical science,
but neither is it a merely practical limitation that we could someday shed. It is an
important fact about the way our intelligence functions in this world, and chaos
theory is interesting because it makes our finitude have real consequences for our
science.66

Recognition of this limitation on predictive ability not only revises our expectations but
enjoins an equal reconsideration of the methodological assumption which underlies its use.
This assumption, previously stated, is that small causes entail as consequences small effects,
and that any anomalous or chaotic behavior resulting from vagueness in measurement can be
compensated for through adjustment and approximation. This assumption is methodological
because it guides scientists in the formation of hypotheses and provides a measure for the
adjudication of competing explanatory schemes. Accordingly, if a system exhibits chaotic

66
​ ​Kellert,​ ​p.​ ​41-2.

105
behavior it is a signal that the theory is wrong or incomplete, because small fluctuations are
not supposed to have widely divergent effects. This leads investigators to alter their theory
and come at it again, but, guided by the same assumption, they encounter the same
difficulties and thus they resort to perturbation analysis and approximations and end by
rapping their heads against this intractable limit. As Peirce might have characterized it, what
is​ ​needed​ ​here​ ​is​ ​a​ ​change​ ​of​ ​habit.

The habit, or inference-guiding principle, is no longer sufficient for the explanation of a


range, of empirical phenomena, Continued use of it is like jumping down the semiotic rabbit
hole, following each sign to its ultimate logical interpretant, and then starting over until the
right interpretant is hit upon. The aims of inquiry require principles of inference which
obviate this and facilitate successful theory choice. Changing the principle is easy enough but
it is supported by metaphysical beliefs which must first be re-evaluated before anything new
is​ ​adopted.

The metaphysical beliefs underpinning our assumption can be loosely defined under the
rubric of determinism. Determinism is the doctrine which conceives of the components of the
universe as linked in a determinate, clock-like manner. Any event in time is derivable from
an initial set of conditions, plus a law or laws which govern the change of the system over
time, such that each moment flows out of the preceding one in an intelligible, mathematically
specifiable way. In addition it conceives of the world's components as intelligibly
self-sufficient units amenable to abstractive analysis from the whole. On a practical level this
belief plays a very comforting role. It assures us that there is an order to the profusion of
stimuli that the world offers and that our inquiries into it are not duped by the hazards of
blind chance nor the designs of some unseen god, benign or malevolent. On the
methodological level, hinging the various strands of determinism, is the notion of
predictability. The kind of world described by determinism provides the conditions for the
full​ ​accounting​ ​of​ ​its​ ​state​ ​at​ ​any​ ​time.

In order to successfully challenge this belief chaos theory must show that one or several
of the elements which characterize it is theoretically impossible to achieve. The elements
composing this doctrine are both metaphysical and epistemological in nature. Chaos theory
has already addressed the epistemological limits of determinism by undermining the hope of
total predictability. The question now concerns the implications for the metaphysical

106
elements. If exact prediction is theoretically impossible what meaningful conclusions can be
drawn​ ​about​ ​what​ ​the​ ​universe​ ​is​ ​like?

Determinism says that systems are smoothly differential, follow a unique evolution as
they change over time, and are composed of variables with precise values. The interesting
thing is that these are exactly the kind of systems chaos theory studies, yet which finds that
they are unpredictable. At the very least, any notion of determinism which includes total
predictability must be false. This, however, is not a potent enough challenge to the doctrine.
John Earman has argued that even though we may not be able to predict the exact trajectory
of a system, that doesn't mean that it does not exist, that the system does not follow a unique
path fixed by preceding conditions and the laws which govern the universe.67 It may be
similarly argued that just because we cannot see atoms, that is no reason to scrap atomic
theory. The difference though is that we believe that atoms exist because their existence have
empirical consequences. The theories which predict them are based on inferences made from
empirical data. This can perhaps be seen as an application of Peirce's pragmatic maxim. Our
conception of something is meaningless if its effects have no practical bearings. The world
may be deterministic or it may be composed of an infinite number of invisible green bunnies.
Peirce says that deciding which is the better hypothesis involves determining how their
consequences manifest themselves in the process of inquiry. If no such consequences are
determinable,​ ​then,​ ​epistemologically,​ ​they​ ​are​ ​equivalent​ ​conceptions.

Certainly determinism recommends itself far more cogently than green bunnies. At the
very least our common Sense experience of the world tells us that determinism is what is
operative in nature. However, this experience is limited to the edge of the "clearing", bound
by the behavior of systems which become turbulent and unpredictable. This is perhaps akin to
the commonsense opinion once held that the sun revolved around the earth. That point of
view was undermined with new experience. Similarly, it is new experience afforded by work
in chaos theory that strongly suggests that certain of our habits of inference be changed. As I
have just mentioned, a conception with no pragmatic value is suspect and should lead us to be
skeptical. What may strongly suggest that determinism insufficiently explains the universe is
the​ ​conjunction​ ​of​ ​both​ ​chaos​ ​theory​ ​and​ ​quantum​ ​theory​.

The hard kernel at the center of determinism is that events are fixed by what has gone

67
​ ​See​ ​John​ ​Earman,​​ ​A​ ​Primer​ ​on​ ​Determinism​,​ ​(Dordrecht:​ ​D.​ ​Reidel,​ ​1986).

107
before, and thus that, though we may unable to predict it, the universe will evolve in one
unique way with no chance of it being otherwise. This means that if two worlds were
conceived with identical initial conditions and as governed by the same law, they would
evolve in exactly the same way. The insights of both chaos theory and quantum theory say
that this is not possible. Quantum theory says that particles cannot be conceived as occupying
a single finitely specifiable point in phase space but rather with a "patch" of area. Recall that
quantum theory is not an assessment of practical limitation but a postulate about the way
things really are. Thus it says that abstracting this patch to an idealized point is not only
practically impossible but theoretically so as well. But even so, a determinist may claim that
it is possible for this patch itself to follow a unique trajectory and thus be determinate with
regard to its future state. It is here that the addition of constraints posed by chaos theory deals
a decisive blow to the determinist conception. According to the sensitive dependence on
initial conditions that chaotic systems display, any two "points" in phase space initially very
close together will evolve on widely differing trajectories. "[T]he common feature of
dynamical systems having a suitably high degree of instability is that each finite region of
state space, no matter how small, contains points that move along rapidly diverging or
qualitatively distinct type of trajectories."68 This means that in a world where chaotic
behavior is at work, the indeterminacies of quantum mechanics provides the conditions
whereby the evolution of more than one trajectory is possible. As Kellert has stated, "Chaos
and quantum theory lead us to just such a vision of the universe as a congeries of interrelated
but open possibilities, foaming forth in its infinitude. Determinism is not so much proven
false​ ​as​ ​rendered​ ​meaningless."69

If this conclusion is right, it is not to say that determinism is not capable of explaining
some range of phenomena, for it certainly does. As I have repeatedly stated, chaos theory
studies systems which, under given conditions, are simple and straight forwardly
deterministic. Where determinism breaks down is in the appearance of chaotic behavior. That
it cannot account for it must lead us to consider skeptically its viability as an explanation for
how the universe as a whole works. Though chaos theory is critical of the determinist view, it
does not on the other hand promote a view of the world as arbitrary and thoroughly random.

68
​ ​Misra,​ ​I.​ ​Prigogine,​ ​and​ ​M.​ ​Courbage,​ ​"From​ ​Deterministic​ ​Dynamics​ ​to​ ​Probabilistic​ ​Description"​ ​Physica​ ​98A:​ ​1-26.
69
​K​ ellert,​ ​p.​ ​74.

108
More than being a negative description of epistemological limits it makes a positive
contribution​ ​to​ ​the​ ​reconception​ ​of​ ​the​ ​metaphysical​ ​assumptions​ ​which​ ​guide​ ​inquiry.

Conclusion

The concluding chapter will be concerned with an assessment of Peirce's metaphysics


and the success he has in recuperating the pragmatic maxim in terms of his thought as a
whole. One of my contentions will be that his metaphysics models quite accurately the new
conceptual paradigm suggested by research in chaos theory. In explaining his metaphysics II
will use what I have discussed here in illustrating some of the more abstract elements of his
thought, and in so doing provide a philosophical perspective from which to understand more
clearly the kind of universe chaos theory suggests. Briefly though, and in general, the kind of
world chaos theory reveals is a holistic one, one whose parts cannot be completely
understood in isolation from one another, for every part affects, to some degree, every other.
And it is a world in which chance and novelty play a real role in the course of evolution. Like
the pruning of a tree to allow for new life, behavior which had hitherto been considered
negatively as senseless and without meaning may be the very means by which nature escapes
the jaws of entropy and fashions new and vital life. This world is not arbitrary and without
order. Strange attractors reveal in their fractal dimensions a world capable of reconciling both
order and randomness in a balance that denies the old metaphor of the clearing/thicket,
suggesting instead a coin with two sides. Understanding this wholeness cannot be done
quantitatively but rather qualitatively. Modeling the fractal dimensions of chaotic systems
allows investigators to understand the properties of the system as a whole, and to characterize
it in terms of global patterns and holistic regularities. Though the world may be
fundamentally unpredictable, at least we now understand how these limits arise in our
experience of the world. Finally, in rather timely accord with disciplines spanning the
humanities, chaos theory recognizes the importance of approaching systems analysis in
historical terms. The state of a system at any one time is not sufficient to determine its future.
Only by considering how the system develops as a whole can any handle on its future
configuration​ ​be​ ​had​ ​at​ ​all.70

70
​ ​For​ ​a​ ​cogent​ ​treatment​ ​of​ ​the​ ​relation​ ​between​ ​chaos​ ​theory​ ​and​ ​literary​ ​studies​ ​see​ ​Hayles,​ ​1990.

109
Compared with Newtonian physics, chaos theory is in its infancy. It will take many
many years of continued research and discussion among the community of inquirers for its
insights to be further confirmed and for it to become an established part of the way science is
done. This is one possible trajectory of its evolution, one I see as likely and fruitful, but one
not without its hazards. It is fruitful not only for scientific investigation but for inquiry into
many different realms of human experience. It is potentially hazardous though if care is not
taken to understand it in its context and with consideration for the limits of inference and
conjecture. Its popularization and co-option by all manner of people out to make a case can
easily blur important distinctions and dilute the important contributions it has to make, as
Peirce all too similarly experienced with the pragmatic maxim. As I move on to conclude my
thesis I will clearly define these limits, but I hope also to provide scope for assessing Peirce's
metaphysics​ ​in​ ​the​ ​context​ ​of​ ​actual​ ​on-going​ ​courses​ ​of​ ​scientific​ ​inquiry.

110
Chapter​ ​Six

The​ ​Metaphysics​ ​of​ ​Evolution

As I remarked earlier, that Peirce devoted time and energy to metaphysical speculation
has been seen by many as an unfortunate lapse into a way of thinking thoroughly at odds with
the concerns of his earlier thought. After all, Peirce himself said that pragmatism shows that
"almost every proposition of ontological metaphysics is either meaningless gibberish . . . or
else is absurd" (5.423). But let the word 'almost be emphasized. Metaphysics is the study and
explanation of the structure of the real, of how the world is. As such, it is a rational object of
inquiry, and so to dismiss it out of hand would be to commit, for Peirce, the greatest sin of the
life of the mind -- blocking the road of inquiry. Whether we like it or not metaphysical
assumptions are an integral part of the work of inquiry and so disregarding them only leads to
a poorer uncritical metaphysics, the propositions of 'almost every' one of which has up until
now been either gibberish or absurd. The pragmatic maxim is a potent and useful tool of
inquiry and should be employed in metaphysical speculation insofar as it clarifies what we
mean, but not such that its reification as a metaphysical principle in itself would preclude
inquiry​ ​from​ ​its​ ​work.

An adequate metaphysics is important if the inferences science makes are to be fruitful


and lead inquiry closer to settled opinion. To this end science has been aided largely by what
Peirce calls instinctual common-sense, or what Galileo before him called ​il lumen naturale.
Peirce contends that the elemental laws of mechanics, laws whose mechanisms are made
manifest in the profusion of our experience of the environment around us, are neither
constructed nor deduced but gleaned, so to speak, owing to the uncanny affinity the knowing
mind has with the environment it knows. The affinity is not so uncanny though. More
abstractly it is explicable in terms of Peirce's realist conception of universals and the critique
of meaning it entails. The real is of the nature of a general and thus that which it means
cannot be reduced to a sense datum. General concepts can only be understood as habits, or as
Apel has remarked, "as a real embodiment of generality understood as a continual regulation
of praxis."71 Apel continues to say that, "Even in this, Peirce's earliest conception of
Pragmatism, there is an approach to metaphysics which permits us to see human habits as

71
​ ​Apel,​ ​p.​ ​143.

111
analogous to, and a continuation of, natural laws."72 It is Peirce's metaphysics which will give
a full accounting of this, necessary because the aforementioned affinity has its practical
limits. The instincts which aided in procuring the basic necessities of life and revealing the
simple laws of mechanics are not adequate to the complex and abstract notions of atomic
physics and electrodynamics. A heuristic "lumen", not a natural one, is needed if science is to
plumb​ ​further​ ​in​ ​the​ ​course​ ​of​ ​inquiry.

Still practicing the lessons he learned from Kant, Peirce approaches metaphysics by
defining its study in terms of a sort of deduction. His speculation is framed by asking, "What
must reality be like given the logic of inquiry and the manner in which reality is intelligible to
us?" The answer he begins to form is plucked not from flights of fancy or imagination but is
grounded, following Kant, in the fruits of his logical investigations. The results of those
investigations pertinent to his concern with metaphysics are two in number: 1) the regulative
principles​ ​of​ ​the​ ​logic​ ​of​ ​inquiry,​ ​and,​ ​2)​ ​his​ ​analysis​ ​of​ ​the​ ​three​ ​fundamental​ ​categories.

Peirce's analysis of the logic of inquiry revealed the supposition of a Real, independent
of what any one person may think of it. Without it inquiry cannot achieve its aims, and so it
must be the purpose of metaphysics to describe that reality, to, as Hookway phrases it, ...
repay the regulative loans taken out in our logical investigations."73 Secondly, his categorial
analysis, which occupies fundamental structural roles in the logic of inquiry, semiotic, and
the normative sciences, will serve Peirce in the conceptual formulation of his metaphysics as
well. His phenomenological analysis of the categories revealed their universal applicability,
as being capable of accounting for all phenomena. If this is true, then it is only reasonable for
metaphysical speculation to be conceptualized in terms of the categories. It is with this brief
prolegomena​ ​to​ ​the​ ​study​ ​of​ ​metaphysics,​ ​if​ ​you​ ​will,​ ​that​ ​Peirce​ ​begins.

His principle writings on metaphysics are contained in a series of papers published in


The Monist: "The Architecture of Theories' (6.7-34), "The Doctrine of Necessity Examined'
(6.35-65), and "The Law of Mind' (6.102-63). The title of the first paper appropriately
introduces the fundamental theme of metaphysics -- structure. Of what is the world formed
and how does it "hang" together? A regular architect uses two by fours and nails in
construction but a philosophical architect uses concepts and ideas, and more fundamentally,

72
​ ​Ibid.
73
​H​ ookway,​ ​p.​ ​264.

112
categories. The notion of law as a material of construction in metaphysics is a fundamental
one,​ ​one​ ​upon​ ​which​ ​Peirce​ ​focuses​ ​his​ ​attention​ ​in​ ​this​ ​paper.

Law,​ ​Spontaneity,​ ​and​ ​the​ ​Structure​ ​of​ ​the​ ​Cosmos

Peirce's question regards the nature of law and its origin. Can law be said to have
always existed, or does the very fact of its existence prompt us to ask questions about it? The
very way in which he asks this question is a consequence of his meaning-critical
transformation of Kant's epistemology. For Kant the thing-in-itself was the ultimate source of
meaning, but unknown and inexplicable. Peirce's principle however postulates the
fundamental cognizability of the Real and thus gives every 'why' question an a priori
answerability. It is a postulate, a regulative hope, but it provides definite scope to the work of
inquiry, making the conception of anything as absolutely brute and inexplicable, nonsense. It
is law which traditionally has this character, of being absolute and the furthest we can go.
Law explains, and behind that we can go no further. Of course for Peirce this constitutes the
sin of blocking the road of inquiry. Of all things, law and regularity seem to call out for
explanation- If, in flipping a coin, heads and tails come up in a random succession, we think
nothing of it. However if heads comes up ninety-five percent of the time we ask why. There
is a regularity there that calls for explanation. Peirce's accounting of law begins with the
following​ ​observation,

Now the only possible way of accounting for the laws of nature and for uniformity
in general is to suppose them the results of evolution. This supposes them not to be
absolute, not to be obeyed precisely. It makes an element of indeterminacy,
spontaneity,​ ​or​ ​absolute​ ​chance​ ​in​ ​nature​ ​(6.13).

Normally, the explanation of something is made by recourse to a law. Thus in treating of the
nature and origin of law itself it would seem, prima facie, to imply the search for more
general laws, continuing in Chinese box fashion ad infinitum. Proceeding by the
presupposition of increasingly general laws gets nowhere in response to the original question
about law itself. Peirce's approach is to treat the existence of laws as a historical
development, as the result of an evolutionary process. His aim in his metaphysics is to
describe the historical conditions which allow for the emergence of law in the first place. His
aim​ ​then,​ ​as​ ​he​ ​says,​ ​is​ ​not​ ​so​ ​much​ ​"a​ ​cosmology,​ ​as​ ​a​ ​cosmogony"​ ​(6.33).

113
In the quote above Peirce introduces the notions of chance and spontaneity as ingredient
in the evolution of law. The origin of this notion in Peirce's thought comes, heuristically,
from the implications of the category of Firstness, but more practically from his sustained and
deep involvement in the scientific developments of his day. Chief among these was the
development of statistical method in the analysis of dynamical systems, most notably in the
work of Darwin and in Maxwell's theory of gases. He regarded the element of chance
variation in Darwin's theory as a vital component in any theory attempting to explain the
variety,​ ​multiplicity,​ ​and​ ​ability​ ​for​ ​growth​ ​that​ ​the​ ​world​ ​displays.​ ​Of​ ​Darwin​ ​he​ ​says,

The theory of Darwin was that evolution had been brought about by the action of
two factors: first, heredity, as a principle making offspring nearly resemble their
parents, while yet giving room for "sporting," or accidental variations . . . and,
second, the destruction of breeds or races that are unable to keep the birth rate up
to​ ​the​ ​death​ ​rate​ ​(6.15).

Darwin's notion of chance-variation certainly influenced the conception of Peirce's


evolutionary theory, but chance for Peirce plays a deeper than merely methodological role.
As​ ​Apel​ ​notes,

. . . Peirce was not content to view statistical theories as a substitute - possibly


temporary -- for a deterministic explanation of individual events, an explanation
which​ ​cannot​ ​be​ ​given​ ​at​ ​the​ ​moment,​ ​owing​ ​to​ ​a​ ​lack​ ​of​ ​knowledge.74

Peirce does not opt for the explanation of chance in statistical terms, as describing the limits
of deterministic laws, but, more radically, deriving lawfulness as a historical development
from chance. As chaos theorists criticize perturbation analysis and attempts at approximation,
so Peirce, while acknowledging the importance of Darwin's insight, sought an understanding
of​ ​chance​ ​apart​ ​from​ ​the​ ​conception​ ​of​ ​the​ ​universe​ ​in​ ​deterministic​ ​terms.

In "The Architecture of Theories" Peirce gives perhaps the most concentrated


exposition​ ​of​ ​his​ ​metaphysics.​ ​As​ ​he​ ​says,

A Cosmogonic Philosophy . . . would suppose that in the beginning -- infinitely


remote -- there was a chaos of unpersonalized feeling, which being without
connection or regularity would properly be without existence. This feeling,
sporting here and there in pure arbitrariness, would have started the germ of a
generalizing tendency. Its other sportings would be evanescent, but this would
have a growing virtue. Thus, the tendency to habit would be started; and from this,
with the other principles of evolution, all the regularities of the universe would be
74
​ ​Apel,​ ​p.​ ​151.

114
evolved. At any time, however, an element of pure chance survives and will
remain until the world becomes an absolutely perfect, rational, and symmetrical
system,​ ​in​ ​which​ ​mind​ ​is​ ​at​ ​last​ ​crystallized​ ​in​ ​the​ ​infinitely​ ​distant​ ​future​ ​(6.33).

The evolutionary quality of Peirce's metaphysics and the novel role that chance play are born
partly of his assessment and rejection of a mechanistic, deterministic model of the universe.
The core of his argument is given in "The Doctrine of Necessity Examined" and is composed
of two themes. The first, as we have seen, is that the existence of regularities requires
explanation. If this is attempted using an increasingly general hierarchy of laws the question
never gets answered. For Peirce it is only a principle of evolution, operating by chance
variation, that eludes this problem. He certainly does not say that chance itself makes
explicable the growth and variety of the world, but only that it provides the opportunity for it
- - "I make use of chance chiefly to make room for a principle of generalization, or tendency
to form habits, which I hold has produced all regularities."75 It is evolution which accounts
for this, a process which needs no explanation because the fortuitous action of chance by
which it operates is not governed by law and is impelled by no outward cause. It exhibits only
a tendency to growth which "can be supposed itself to have grown from an infinitesimal germ
accidentally​ ​started"​ ​(6.14).

The second prong of his argument concerns the heterogeneous variety of the observed
world. A process which is mechanistic, Peirce says, is characterized by reversibility, a fact
which renders inexplicable the phenomena of novelty, growth and development. Adducing
examples from the plant and animal world, from the analysis of mind, of institutions,
language, and ideas, Peirce infers that "there is probably in nature some agency by which the
complexity and diversity of things can be increased; and that consequently the rule of
mechanical​ ​necessity​ ​meets​ ​in​ ​some​ ​way​ ​with​ ​interference."​ ​He​ ​continues,

By thus admitting pure spontaneity or life as a character of the universe, acting


always and everywhere though restrained within narrow bounds by law, producing
infinitesimal departures from law continually, and great ones with infinite
infrequency,​ ​I​ ​account​ ​for​ ​all​ ​the​ ​variety​ ​and​ ​diversity​ ​of​ ​the​ ​universe​ ​.​ ​.​ ​.​ ​.76

75
​ ​Philip​ ​P.​ ​Wiener,​ ​ed.,​ ​Charles​ ​S.​ ​Peirce:​ ​Selected​ ​Writings​,​ ​(New​ ​York:​ ​Dover​ ​Publications,​ ​1958),​ ​p.​ ​177.
76
​ I​ bid.,​ ​p.​ ​174-5.

115
Returning then to the extended quote from "The Architecture of Theories," Peirce
conceives the lawfulness of the universe as having begun in a "soup" of pure arbitrariness and
indeterminacy - - of chaos it may be said. Using the word 'sportings' he attempts to describe
the slow and fortuitous formation of a generalizing tendency within this state such that
existents are ushered into being which then enter eventually and over time into relational
groupings. This tendency to interaction signals the formation of habit, of regularity, or of
what​ ​was​ ​equivalent​ ​for​ ​Peirce,​ ​law.

The general movement here is clearly one from arbitrariness and chaos to order and law,
and in fact Peirce seems logically to conclude the passage by referring to the outcome of such
a process as "an absolutely perfect, rational, and symmetrical system, in which mind is at last
crystallized in the infinitely distant future" (6.33). The conclusion drawn here seems destined
from his earliest thoughts on the nature of inquiry, that opinion will converge in the long run
and that with the rigidification of habits of inference, inquiry will come to rest. He thus
seems, in Hegelian fashion, to reduce his first and second categories to a process of being
eventually overcome and subsumed by the mediating and lawlike category of Thirdness. He
in no way intends this however. The future point he describes is an infinitely future one,
acting purely regulatively as an agent in the logic of inquiry. The influence of chance
variation continues as a real component of the dynamics of the universe, acting as a goad to
the work of evolution. Furthering this end, Peirce comments on the possibility of exact
measurement​ ​saying,

[O]bservations which are generally adduced in favour of mechanical causation


simply prove that there is an element of regularity in nature, and have no bearing
whatever upon the question of whether such regularity is exact and universal or
not. Nay, in regard to this exactitude, all observation is directly opposed to it. . . .
Try to verify any law of nature, and you will find that the more precise your
observations, the more certain they will be to show irregular departures from the
law. We are accustomed to ascribe these . . . to errors of observation; yet we
cannot usually account for such errors in any antecedently probable way. Trace
their causes back far enough and you will be forced to admit they are always due
to​ ​arbitrary​ ​determination,​ ​or​ ​chance​ ​(6.46).

This inability to ascribe exact values to observational phenomena imply that the future point
of convergence characterized in the logic of inquiry is never fully attained but only

116
asymptotically approached. The divergence of opinion will decrease, its fluctuations coming
to​ ​hover​ ​within​ ​a​ ​certain​ ​range,​ ​but​ ​not​ ​such​ ​that​ ​it​ ​will​ ​ever​ ​reach​ ​an​ ​exact​ ​value.

Thus Peirce conceives of chance not as some statistical vagueness to be overcome in the
evolution of law, but as a real constituent in the way the universe works. The determinist
views chance as anomalous and inexplicable because his or her metaphysics is insufficient in
scope. Peirce says, "We may, therefore, say that a world of chance is simply our actual world
viewed from the standpoint of an animal at the very vanishing-point of intelligence. The
actual​ ​world​ ​is​ ​almost​ ​a​ ​chance-medley​ ​to​ ​the​ ​mind​ ​of​ ​a​ ​polyp."​ ​(6.406).

Evolution​ ​and​ ​the​ ​Categories:​ ​Tychism,​ ​Anancasm,​ ​and​ ​Agapism

The name Peirce gives to his theory of chance-spontaneity in evolution is "tychism,"


which can be understood as a direct specification of his theory of the categories. Tychism
illustrates the category of Firstness, connoted by the undifferentiated continuum of
possibilities inherent in the universe, especially in the initial stages of evolution, possibilities
realized by chance variation and spontaneity. The character of Secondness in the process of
evolution Peirce terms "anancasm" and is illustrated by the familiar Darwinian notions of
struggle, reaction, and "survival of the fittest." It is this aspect of the evolutionary dynamic
which most readily displays the mechanistic principles with which we are familiar and to
which Peirce understands the limit of determinism's scope. To be clear, Peirce is not
describing two different theories of evolution nor two different outcomes of the process, but
the very self-same process from two points of view. His categorial theory maintains that
phenomena of every kind exhibit Firstness, Secondness, and Thirdness, and so it is by this
heuristic​ ​that​ ​his​ ​speculation​ ​is​ ​guided.

Accordingly, he analyzes the evolutionary process in terms dictated by the category of


Thirdness. His exposition is contained in his articles "The Law of Mind" and "Evolutionary
Love" in which he conceives of what he calls "agapism." Having been influenced by Darwin
in the conception of tychism and anancasm, he turns, perhaps oddly to the modern reader, to
Lamarck. Darwin's theory turned on the notion of chance variation, its results being
propagated to future generations as an inherited characteristic. Lamarck's theory attempted to
understand the process of evolution as a striving in which characteristics are acquired. It is
the idea of purposiveness which characterizes Lamarck's theory, as opposed to a blind and

117
merely random process, and is the correlate for Peirce's discussion of the growth of law and
reasonableness​ ​in​ ​the​ ​cosmos.

In​ ​discussing​ ​the​ ​idea​ ​of​ ​spontaneity​ ​Peirce​ ​says,

But my hypothesis of spontaneity does explain irregularity, in a certain sense; that


is, it explains the general fact of irregularity, though not, of course, what each
lawless event is to be. At the same time, by thus loosening the bond of necessity, it
gives room for the influence of another kind of causation, such as seems to be
operative​ ​in​ ​the​ ​mind​ ​in​ ​the​ ​formation​ ​of​ ​associations​ ​(6.60).

As a response to determinism Peirce offers agapism as this other kind of causation. The term
he coins here comes, of course, from the Greek agape, meaning love, or more pointedly in
this context, her names attraction, and he means to denote by it the tendency toward
generalization and habit-taking, the formation of patterns and regularities in consciousness
and​ ​in​ ​nature.​ ​He​ ​says,

Logical analysis applied to mental phenomena shows that there is but one law of
mind, namely, that ideas tend to spread continuously and to affect certain others
which stand to them in a peculiar relation of affectability. In this spreading they
lose intensity, and especially the power of affecting others, but gain generality and
become​ ​welded​ ​with​ ​other​ ​ideas.77

It could be asked why this tendency toward generalization could not be understood
anancastically, that is, mechanically. Peirce responds by saying that the progression of a
deterministic universe is inevitable and necessary. It lacks purposiveness because its
teleology is ineluctably compelled. On the contrary he characterizes his view of agapism as a
developmental teleology, in that the occurrence of chance variation can create new purposes
and​ ​ends.​ ​He​ ​likens​ ​this​ ​notion​ ​to​ ​the​ ​phenomenon​ ​of​ ​human​ ​personality,

. . . in the case of personality this teleology agapism) is more than a mere


purposive pursuit of a predeterminate end; it is a developmental teleology. ... Were
the end of a person already explicit, there would be no room for development or
growth, for life. . . . The mere carrying out of predetermined ends is mechanical.
(6.156-7).

77
​ ​Peirce,​ ​C.​ ​S.,​ ​The​ ​Philosophical​ ​Writings​ ​of​ ​Peirce​ ​.​ ​(New​ ​York:​ ​Dover​ ​Publications:​ ​1955),​ ​p.​ ​340.

118
Hausman has characterized Peirce's understanding of purpose in this sense as
"self-determining action."78 This is aptly phrased, for though agapism assumes and
incorporates tychistic and an anancastic elements, it is not reducible to them, in the same way
that Peirce felt that the pragmatic maxim was not reducible to mere Secondness or action.
Though tychism is an aspect of his evolutionary metaphysics, Peirce says that, "it only enters
as subsidiary to that which is really, as I regard it, the characteristic of my doctrine, namely
that I chiefly insist upon continuity or Thirdness" (6.202). This generalizing tendency is
neither heedlessly arbitrary nor determinantly mechanistic but rather creatively dynamic. The
association and growth of ideas and habits is a process compelled not by some external
overarching law but fueled rather by a holistic dynamism as it responds and adapts to
instances of chance spontaneity. Evolution as agapism is "self-determining" then in the sense
that, since the development of the cosmos has no predetermined telos, it must be open to the
departure from and creation of new habits or laws. The development that ensues is nowhere
recorded in the minutiae of some law but is a future whose dimensions are seen only broadly,
measured by the limit of genuine openness and creativity to which the world is prone, a limit
which​ ​for​ ​Peirce​ ​is​ ​the​ ​positive​ ​condition​ ​for​ ​growth,​ ​variation,​ ​and​ ​life.

The notion of habit, or habit-taking, is of paramount importance to Peirce in his


metaphysics, occupying that space generally filled by the notion of law. Clearly the bulk of
his writing characterizes law as exemplifying the category of thirdness. Considered as a mode
of being, for example, law is conceived of as a Third. But in his metaphysics Peirce is
concerned with a slightly different aspect of law, that is, with law understood as a mode, not
of being, but of existence. He says, "Law as an active force is second but order and
legislation are third" (1.337). It is in terms of the actually operative force of law that Peirce
theorizes its relation with chance, and thus discerns its contingent character. Thus, though we
have considered Peirce's notions of Tychism, Ananasm, and Agapism, we can alternatively
understand the categorial conception of his metaphysics in terms of the triad of chance, law,
and habit, perhaps thereby making clearer the relation he understands between law and
spontaneity.

Chance, as we have seen, is his first category. Chance, he says, "or irregularity -- that is,
the absence of any coincidence, is that diversity and variety of things and events which law

78
​ ​Carl​ ​R.​ ​Hausman,​ ​Charles​ ​S.​ ​Peirce's​ ​Evolutionary​ ​Philosopohy​.​ ​(Cambridge:​ ​Cambridge​ ​University​ ​Press,​ ​1993),​ ​p.​ ​176.

119
does not prevent" (6.612). Ontologically, chance is an existential specification of the broad
conception of firstness, that is, as pure qualitative possibility. It is a real constituent of the
universe and not due, as we have seen, to any sort of subjective ignorance. It is sui generis
and is as it is without relation to any other thing. As a fortuitously occuring event then,
chance​ ​is​ ​not​ ​something​ ​connected​ ​or​ ​determined​ ​by​ ​law.

Law is the second category of existence for Peirce, and is that force by which future
events are governed. Law, as Peirce characterizes it here, seems to have the connotation of
efficient, or mechanical causation, being that calculus or mechanism by which events follow
one another with regularity. As we have seen, however, these laws are not obeyed precisely.
James​ ​Feibleman

describes​ ​and​ ​explains​ ​this​ ​in​ ​the​ ​following,

The discrepancy between the absolute generality which we might suppose law to
have and the limited generality of actual law is accounted for by the fact that law is
not a mere uniformity but a compulsion. "Let a law of nature - - say the law of
gravitation -- remain a mere uniformity ... and what in the world would induce a
stone​ ​.​ ​.​ ​.​ ​to​ ​act​ ​in​ ​conformity​ ​to​ ​that​ ​uniformity?"​ ​(5.48).79

The question is, what compels this uniformity? The answer lies in the notion of habit. As
Peirce says, since "a law is how an endless future must continue to be," it follows that "a law
never​ ​can​ ​be​ ​embodied​ ​in​ ​its​ ​character​ ​as​ ​a​ ​law​ ​except​ ​by​ ​determining​ ​a​ ​habit"​ ​(l.536).

Habit is the third category of existence for Peirce, or more pointedly, habit-taking, for
habit, as such, connotes merely the uniformity mentioned above. For Peirce, when "accident
acquires some incipient staying quality, some tendency toward consistency," then "some
beginning of a habit has been established" (6.204). He gives a straightforward example in the
way a stream of water forms a bed for itself. The continued uniformity of events grows or
"carves out" a habit, but it is the ability itself, the habit-taking, to which he refers, and which
corresponds with the notions he describes in Agapism, that is, the generalizing growth of
ideas.

However, it is chance-spontaneity which provides the formation of habit with its sphere
of influence. Taken together it is the element of chance and the growth of habit-taking which
Peirce​ ​understands​ ​as​ ​accounting​ ​for​ ​law.​ ​As​ ​Peirce​ ​says,

79
​ ​James​ ​K.​ ​Feibleman,​ ​An​ ​Introduction​ ​to​ ​the​ ​Philosophy​ ​of​ ​Charles​ ​S.​ ​Peirce​,​ ​(Cambridge:​ ​MIT​ ​Press,​ ​1970),​ ​p.​ ​192.

120
... if law is a result of evolution . . . it follows that no law is absolute. That is, we must
suppose that the phenomena themselves involve departures from law analogous to
errors of observation (6.101). Yet) habits produce statistical uniformities and when
these become great enough there are at least no departures from the law that our
senses can take cognizance of. It is clear that nothing but a principle of habit, itself due
to the growth by habit of an infinitesimal chance tendency toward habit-taking, is the
only bridge that can span the chasm between the chance medley of chaos and the
cosmos​ ​of​ ​order​ ​and​ ​law​ ​(6.262).

Given this overall scheme we can see Peirce's conception of evolution describing a
general movement from an initial prominence of arbitrariness to one of increasing order and
law. The tendency toward generalization is the operative dynamic and principle intelligible
structure of the cosmos but it would be empty and merely mechanical if chance were not a
real element in evolution. Though law and regularity predominate in the universe they are not
inviolate and eternal. The growth of Thirdness is possible because the Firstness of the
universe​ ​is​ ​never​ ​overcome​ ​and​ ​fully​ ​rationalized.​ ​As​ ​Peirce​ ​said,

Thus these two elements, at least, exist in nature, Spontaneity and Law. Now, to
ask that spontaneity should be explained is illogical, and indeed absurd. But to
explain a thing is to show how it may have been a result of something else. Law,
then, ought to be explained as a result of Spontaneity. Now the only way to do that
is to show in some way that law may have been a product of growth, of evolution.
80

By this analysis Peirce reconceives the notion of law in some middle ground between the
extremes of pure chance and thorough order. Agapism is a hypothesis which combines these
extremes in such a way as to make room (on the one hand) "for a principle of the tendency to
form​ ​habits,​ ​or​ ​(on​ ​the​ ​other​ ​hand)​ ​for​ ​a​ ​kind​ ​of​ ​spontaneity​ ​that​ ​is​ ​to​ ​some​ ​degree​ ​regular."81

80
​ ​Richard​ ​S.​ ​Robin,​ ​Annotated​ ​Catalogue​ ​of​ ​Charles​ ​Sanders​ ​Peirce​,​ ​(Ameherst:​ ​University​ ​of​ ​Massachusetts​ ​Press,​ ​1967),
950.00010-1.
81
​ ​Peirce,​ ​as​ ​quoted​ ​by​ ​Hausman,​ ​p.​ ​173.

121
Peirce​ ​and​ ​Chaos​ ​Theory

Discussing the speculative and highly abstract determinations of the categories in Peirce's
metaphysics can lead one, before long, into feeling swamped by a certain vagueness
originating not only from the difficult and abstract nature of the subject matter but from the
very words Peirce coins to describe it. It is at this point that I want to return to chaos theory in
the​ ​hope​ ​of​ ​providing​ ​concrete​ ​illustration​ ​of​ ​Peirce's​ ​theory.

A first line of comparative interpretation is provided by the similar critique of


determinism which informs the work of both Peirce's theory and chaos theory. As I quoted
Peirce earlier, "Try to verify any law of nature, and you will find that the more precise your
observations, the more certain they will be to show irregular departures from the law" (6.46).
Methodologically, both theories regard determinist principles as efficacious in the
explanation of a limited range of phenomena, but beyond that, the inadequacy of determinism
to explain phenomena such as chaotic systems for example, becomes reflected in the
metaphysical assumptions of both Peirce and chaos theory, assumptions which make chance
and chaos something more than unintelligible anomalies, and which thus makes determinism
an inadequate theory of the universe as a whole. Chaos theory abjures the possibility of
precise measurement, and thus prediction, because the real decimal expression of any
variable is potentially infinite. It is in that part that doesn't get measured, in the inherent
"vagueness" of every measurement, that chaotic behavior arises, behavior which traditional
methods are incapable of grasping. Beneath, as it were, the range of predictable systems, lies
what chaos theory deems a more fundamental dynamic of the universe, one composed of a
holistic interaction and dependence between all levels of dynamical systems in the cosmos
and characterized by a sensitivity to initial conditions which gives to chance and
indeterminacy​ ​a​ ​real​ ​constitutive​ ​role​ ​in​ ​the​ ​formation​ ​of​ ​the​ ​order​ ​we​ ​discover​ ​in​ ​nature.

Peirce's estimation is entirely of a piece with this. Regarding irregular departures from
law​ ​he​ ​says,

We are accustomed to ascribe these . . . to errors of observation; yet we cannot


usually account for such errors in an antecedently probable way. Trace their causes
back far enough and you will be forced to admit they are always due to arbitrary
determination,​ ​or​ ​chance​ ​(6.46).

122
The holistic characterization of the cosmos in chaos theory is akin to Peirce's insistence on
continuity, his denial of atomized discrete units whose sum makes up the content of the
universe. By breaking down the subject/object, mind/matter dualism of Descartes he made
room for the conception of the cosmos as a continuous whole whose various aspects he
discerns in terms of his categories. The primordial character of the world is Firstness, the
pure realm of qualitative possibility, which for illustrative purposes we can envision as pure
undifferentiated energy. The specification of this energy over the course of evolution into
Secondness and Thirdness, that is, existents and the laws which govern them, is the manner in
which Peirce accounts for the formation of law and the variety and diversity of the world we
experience. However the categories are not opposed and wholly other in their relation. Where
for Descartes a rock and the mind which knows it are thoroughly different things, for Peirce
they are merely varying degrees of the perfuse continuum of the cosmos. He called matter
"effete mind" by which he meant that the habits, which define rockness for example, have
become so "hidebound" that they've lost their ability to change and adapt. But they are not
fundamentally​ ​different​ ​things.​ ​As​ ​he​ ​says,

We must . . . regard matter as mind whose habits have become fixed so as to lose
the powers of forming them and losing them, while mind is to be regarded as a
chemical genus of extreme complexity and instability. It has acquired in a
remarkable​ ​degree​ ​a​ ​habit​ ​of​ ​taking​ ​and​ ​laying​ ​aside​ ​habits​ ​(6.101).

The fundamental sameness and connectedness then of the phenomena of the cosmos is
partly what shapes the course of Peirce's thinking about law. Law arises as a result of
evolution, a process which is spurred by the fortuitous events of chance variation. Trace the
evolution of a law back far enough and this is where you end up says Peirce. For him it is
explicated in terms of Firstness. For chaos theory the phenomenon of chaotic behavior is
traceable to the sensitive dependence on initial conditions inherent in nonlinear systems.
Recalling Edward Lorenz's experiment, the shift in measure from three to just six decimal
places produced predictions widely divergent from one another. This calls into question the
determinist assumption that there is a proportionate relationship between cause and effect. On
the contrary, very small causes can have causes of potentially much greater magnitude, which
is to reinforce the continuously holistic dynamic of the world as a whole. Peirce touched on
just this phenomenon many decades before the advent of high speed computers when, in
discussing​ ​the​ ​intensification​ ​of​ ​feeling​ ​in​ ​protoplasm​ ​he​ ​says,

123
. . . habits are general ways of behavior which are associated with the removal of
stimuli. But when the expected removal of the stimulus fails to occur, the
excitation continues and increases, and non-habitual reactions take place; and
these tend to weaken the habit. If, then, we suppose that matter never does obey its
ideal laws with absolute precision, but that there are almost insensible fortuitous
departures from regularity, these will produce, in general, equally minute effects.
But protoplasm is in an excessively unstable condition; and it is the characteristic
of unstable equilibrium that near that point ​excessively minute causes may
produces startlingly large effects.​ Here, then, the usual departures from regularity
will be followed by others that are very great; and the large fortuitous departures
from law so produced will tend still further to break up the laws, supposing that
these are of the nature of habits. Now, this breaking up of habit and renewed
fortuitous spontaneity will, according to the law of mind, be accompanied by an
intensification​ ​of​ ​feeling​ ​(emphasis​ ​mine,​ ​6.264).

Feeling is the anthropomorphic word Peirce uses to refer to the Firstness of the universe, its
pure qualitative possibility or potentiality (much like Aristotle). He speaks of it here in the
terms of stimulus/reaction events of a protoplasm because protoplasm models highly
accurately, in his opinion, the extremely complex and unstable chemical nature of the mind.
The regularity and predictability of law is stable in nature but does not constitute a
hegemony. Chance-spontaneity in the continuum of feeling is an event whose intensification
can potentially disrupt the regularity of habits or laws, producing effects largely
disproportionate​ ​to​ ​those​ ​expected​ ​by​ ​determinism.

It is this very breaking up of predictable habits that is the object of study for chaos
theorists. The onset of chaotic behavior may very well be understood, in Peircean terms, as
the constant, "low-grade" interaction between the Firstness and Thirdness of the cosmos,
between chaos and order. I say "low-grade" because clearly our experience of the world is
largely one of order and predictability, but laws are not static entities. They grow and adapt in
the evolving dynamic of the universe and this growth is possible only if there is a "pruning,"
that is, if chance is a real component of that dynamic. Chaotic behavior, then, as chaos
theorists understand it and as Peirce's speculation seems to suggest, is not an anomalous
aberration but rather the locus of the dynamic growth and evolution of law. As such, though it
is not predictable, it is rational. As Peirce said, " . . . my hypothesis of spontaneity does
explain irregularity, in a certain sense; that is, it explains the general fact of irregularity,
though​ ​not,​ ​of​ ​course,​ ​what​ ​each​ ​lawless​ ​event​ ​is​ ​to​ ​be"​ ​(6.60).

124
Although Peirce and chaos theory share very similar views with regard to determinism
and the events of deviation from the predictions of law, some clarification regarding the
connection I see between the two needs to be made. Though Peirce was a practicing scientist
all his life, he was writing as a philosopher in his speculations about metaphysics. From his
own scientific work as a chemist and astronomer he was aware of the discrepancies to be
found in measurement and thus the manner in which laws are not obeyed precisely, but he
lacked the technical analysis to make scientifically theoretical sense of this. Instead he used
the​ ​language​ ​of​ ​philosophy.​ ​In​ ​a​ ​telling​ ​passage​ ​he​ ​says,

The hypothesis of chance-spontaneity is one whose inevitable consequences are


capable of being traced out with mathematical precision into considerable detail.
Much of this I have done and find the consequences to agree with observed facts
to an extent which seems to me remarkable. But the matter and methods of
reasoning are novel, and I have no right to promise that other mathematicians shall
find my deductions as satisfactory as I myself do, so that the strongest reason for
my belief must for the present remain a private reason of my own, and cannot
influence others. I mention it to explain my own position; and partly to indicate to
future​ ​mathematical​ ​speculators​ ​a​ ​veritable​ ​gold​ ​mine​ ​.​ ​.​ ​.82

In thus drawing comparison between his conception of chance and the array of notions in
chaos theory concerned with the onset of chaotic behavior, I am looking not for an exact
technical correspondence but rather for a philosophical alliance by which the thought of each
can be fruitfully enhanced by the other. In my estimation, the technical descriptions
employed in chaos theory (sensitive dependence on initial conditions, etc.) offer the
mechanism for chance variation that Peirce hypothesized, and in some sense can be seen as
the motherload which Peirce refers to above. Peirce left sketchy hints as to the whereabouts
of​ ​a​ ​gold​ ​mine​ ​and​ ​it​ ​may​ ​be​ ​that​ ​the​ ​work​ ​of​ ​chaos​ ​theory​ ​reflects​ ​some​ ​of​ ​that​ ​very​ ​gold.

Above I say that chaos theory conceives of chaotic behavior as the locus for the
dynamic evolution of law. This may seem in some ways misleading, for clearly the concern
of chaos theorists is with understanding chaotic behavior, whereas the emphasis for Peirce is
on law and its evolution. Peirce insisted on characterizing his overall conception of the
evolution​ ​of​ ​the​ ​universe​ ​as​ ​agapistic.​ ​He​ ​said,

I object to having my metaphysical system as a whole called Tychism. For


although tychism does enter into it, it only enters as subsidiary to that which is

82
​ ​Wiener,​ ​p.​ ​77.

125
really, as I regard it, the characteristic of my doctrine, namely that I chiefly insist
upon​ ​continuity​ ​or​ ​Thirdness​ ​(6.202).

Although Peirce and chaos theory clearly emphasize different concerns, some chaos theorists
have recognized the potential metaphysical implications of chaotic behavior in such a way as
to​ ​appreciate​ ​the​ ​evolving​ ​notion​ ​of​ ​law​ ​which​ ​Peirce​ ​characterizes.

In his book on chaos theory James Gleick describes the thinking of Otto Rössler, the
discoverer​ ​of​ ​a​ ​chaotic​ ​attractor,​ ​saying​ ​that​ ​he,

felt that these shapes fractal attractors) embodied a self-organizing principle in the
world. He would imagine something like a wind sock on an airfield, "an open hose
with a hole in the end, and the wind forces its way in," he said. "Then the wind is
trapped. Against its will, energy is doing something productive, like the devil in
medieval history. The principle is that nature does something against its own will
and,​ ​by​ ​self-entanglement,​ ​produces​ ​beauty.83

In a world conceived as deterministically Newtonian where law operates mechanically, the


second law of thermodynamics would seem to predict a continually degenerating universe.
But clearly this is not the case. Owing largely to Darwin we are aware of the increase in
complexity, organization, and productivity of the world around us. What chaos theory may
provide is the underlying dynamic, the "creative engine" if you will, driving evolution. As
Paul​ ​Davies​ ​says,

The emerging paradigm, by contrast, recognizes that the collective and holistic
properties of physical systems can display new and unforseen modes of behavior
that are not captured by the Newtonian and thermodynamic approaches. There
arises the possibility of self-organization, in which systems suddenly and
spontaneously leap into more elaborate forms. These forms are characterized by
greater complexity, by cooperative behavior and global coherence, by the
appearance of spatial patterns and temporal rhythms, and by the general
unpredictability​ ​of​ ​their​ ​final​ ​forms.​ 84

It must be stressed, however, that currently, the empirical results of chaos theory do not
postulate the evolution of new laws to replace those laws which fail to predict chaotic
behavior. The thoughts above are speculative and though they don't reflect the bulk of
research done in chaos theory, I do believe them to be pertinent indications of the further

83
​ ​Gleick,​ ​p.​ ​142.
84
​P​ aul​ ​Davies,​ ​The​ ​Cosmic​ ​Blueprint​,​ ​(New​ ​York:​ ​Simon​ ​and​ ​Schuster,​ ​1988),​ ​p.​ ​197-98.

126
fruitfulness​ ​of​ ​Peirce's​ ​insights.

Perhaps Peirce's notion of the evolution of law can be tempered a bit here. When he
says that chance-spontaneity allows for the evolution of law, he does not mean that a
particular law (governing a system from which a deviation occurred) is bad or ineffective, but
merely that a new rift is formed in which new uniformities can begin to carve out a habit of
regularity. Given a span of time great enough, the formation of these new habits, via the
generalizing force of Agapism, will have become attracted to other uniformities in such a
way as to have effected an evolution, small though it may be, in the overall "shape" of the
dynamical systems governing the universe. Though such a scenario remains only a
speculative hypothesis, it is not precluded or flatly contradicted by current research in chaos
theory,​ ​and​ ​may​ ​in​ ​fact,​ ​with​ ​refined​ ​analysis,​ ​be​ ​the​ ​object​ ​of​ ​future​ ​discovery.

To clarify thus far, my principal concern with Peirce and chaos theory is to elaborate the
very similar ways in which they each approach the nature of dynamical systems. Both Peirce
and chaos theory note that there are apparently anomalous deviations in dynamical systems
unaccounted for by classically deterministic laws. Both see that laws are not obeyed precisely
and that exact measurement is impossible. The manner in which both begin to account for
this is by trying to understand the role of chance variation. Both stress that the spontaneous
eruption of unpredicted behavior is not the statistical limit of a deterministic law, but rather is
of a nature completely ungoverned by law.85 According to his category of Firstness Peirce
called this 'chance' or 'spontaneity'. This was his hypothesis of some agency by which
deviation from law occurs, and thus by which regularities change and develop. Chaos theory
calls this deviation from law "chaotic behavior' and in its attempt to account for it, for this
agency as Peirce calls it, has elaborated the finely detailed notions of sensitive dependence on
initial conditions, iterative feedback, and fractal attractors - the theoretical mechanisms by
which chance in the universe plays its role. Though Peirce speaks of spontaneity in the most
general of philosophical terms, it seems clear that it is something very much like this
phenomenon​ ​with​ ​which​ ​chaos​ ​theory​ ​has​ ​been​ ​concerned​ ​to​ ​understand.

85
​There is an ambiguity in the use of such terms as "deviation from law" or "behavior ungoverned by law" which should be
clarified. Though a deterministic law cannot predict with precision the outcome of a nonlinear system, it can predict the
general fact and conditions of the chaotic behavior that arises. By using such words as "deviation" and "ungoverned" I mean
to indicate the impossibility of exact measurement and prediction, and thus the alternative methods chaos theory resorts to in
modelling​ ​the​ ​behavior.

127
To quote Peirce again, his tychism, "explains the general fact of irregularity, though not,
of course, what each lawless event is to be." It is with this statement that I move to what I
consider the truly illustrative part of my exposition of Peirce via chaos theory, the notion of
the strange attractor. The statement above can be construed as a qualitative rather than
quantitative formulation, providing a general explanation rather than a precise one, or a
gestalt, if you will, rather than a calculus. "What each lawless event is to be" is a
determination impossible to achieve. It is with this insight that chaos theorists approach an
understanding of chaotic behavior. The phenomenon of sensitive dependence on initial
conditions provides the theoretical basis for understanding the onset of chaotic behavior, but
it is the modeling of it on a strange attractor that now allows us to see it as more than aberrant
white noise. It was perhaps just such a development that W. B. Gallie, writing in 1952,
envisioned as providing future students of philosophy with the tools necessary to appreciate
Peirce's​ ​hypotheses.​ ​He​ ​says,

It seems reasonable to assume, granted continued scientific progress and


intelligent general discussion of scientific results, that in a few decades the kind of
distinction made by Peirce between laws governing reversible and laws governing
irreversible processes will have been both generalized and clarified to cover what
are at present obscure or border-line cases. It would then be possible for us to give
a useful statement of the distinction which we all vaguely recognize, between
those sciences whose laws are primarily (if not exclusively) of a forecasting or
predictive character - sciences which we might describe as 'nomic' -- and those
sciences whose laws serve primarily, not to make predictions, but to unify or
'thicken' our conceptions of different strands of cosmic or terrestrial or biological
or human history - - sciences which we might describe as '-gonic' rather than
'nomic'. Should this supposition prove true, then it will be much easier for future
students of philosophy than it is for us to appreciate the value of Peirce's
cosmology.​ 86

The distinction Peirce hypothesized has been realized in the work of chaos theory and thus it
is​ ​with​ ​the​ ​strange​ ​attractor​ ​that​ ​I​ ​want​ ​to​ ​illustrate​ ​the​ ​kind​ ​of​ ​universe​ ​Peirce​ ​envisions.

As should be clear by now, the actual convergence of opinion in the long run serves
Peirce's logic of inquiry only as a regulative ideal. It was only upon such a hope that Peirce
could conceive of inquiry achieving its aims. However, what is this aim if it is not the
settlement of opinion, the complete settlement of opinion? If this point is never reached then
how are inquiry's aims achieved? If an ​unactualizable ​settlement of opinion is the aim, what,

86
​ ​Gallie,​ ​p.​ ​238.

128
in the final analysis is actually achieved? Or perhaps more pointedly, what must the nature of
the Real, the 'final logical interpretant' that inquiry investigates, be like if the actualization of
such​ ​an​ ​aim​ ​is​ ​ruled​ ​out​ ​of​ ​consideration?

What Peirce has told us so far is that this real consists of at least two elements,
Spontaneity and Law, interacting within an evolutionary matrix of growth and development.
Law, or the generalizing tendency, of itself implies a teleological movement toward
perfection, toward the dominance of Thirdness in the universe. But the categorial reality of
Spontaneity conditions this movement, making of law something never precisely configured
but rather only approximately so. Thus the laws that science seeks to uncover in investigation
are not static entities, permanently ensconced in the shape of the cosmos. Though largely
stable, they are dynamic and always evolving by some small degree; and they will not
crystallize like a diamond at some future point. Such a point was for Peirce lifeless, a state
incompatible with the vitality and energy of the cosmos. The reason he characterized the real
with the final opinion of the community of inquirers was because he saw human habits of
inference as coextensive with the habits or laws by which the universe works. In a very real
sense, following Aristotle, we are that which we know. To hypothesize then an absolute end
to inquiry would be to envision, not only the complete rationalization of the universe, but an
end​ ​to​ ​the​ ​human​ ​condition.

Thus the real is a notion which must answer to the conditions of life and growth as
conceived in Peirce's metaphysics. It must be intelligible, which means it has to be general,
but it can be so only approximately. That there will always be an unintelligible element is the
price paid for the growth and development of the universe. What his notion of the real does
then is to combine chance and order in a delicate but harmonious relation that, on the one
hand, offers inquiry something more determinate than blind chance by which to guide its
course, and on the other, provides the conditions whereby spontaneity is able to sustain the
vitality of the cosmos. If we were to ask, like researchers in systems dynamics, what shape
the cosmos as a system fills in phase space, it is my belief that the answer would be
something​ ​like​ ​a​ ​strange​ ​attractor.

To recall, the shape of a strange attractor is what chaos theorists call a fractal. Its
peculiar configuration, falling somewhere between two and three dimensions, and its unique
scaling properties, make it capable of fitting infinite trajectories within a finite space. Though

129
trajectories never repeat themselves on a strange attractor, they are nonetheless attracted to its
particular shape. It gives investigators a qualitative way of understanding chaotic behavior, of
seeing order in what seems to be only random and arbitrary. In the same way we can
conceive of the collection of dynamical laws which constitutes Peirce's vision of the cosmos.
Within a limited scope and with a certain range of effectiveness, various laws can be
construed deterministically, meaning that highly accurate predictions can be made from them.
However, considering the cosmos as a whole, the best way to account for the irregularities of
otherwise deterministically lawful systems is to conceive of the dynamic between spontaneity
and law as describing the fractal shape of a strange attractor. By this hypothesis, as Peirce
said, the general fact of irregularity is explained, "though not, of course, what each lawless
event is to be" (6.60). Our predictive abilities can plumb the laws of the universe only so far.
Beyond that we are faced with the everactive, creative energies of the universe itself as it
grows and expands. However, with the insights of Peirce and chaos theory, the chaos or
unintelligibility that inheres in the structure of the universe is understood in a new way. Not
only is it theoretically explicable, but it is practically comprehensible. The strange attractor
gives investigators qualitative information about systems as a whole and, as I have
mentioned, increased understanding about the principles governing the onset of chaotic
behavior​ ​has​ ​provided​ ​strategies​ ​for​ ​controlling​ ​chaos​ ​for​ ​the​ ​benefit​ ​of​ ​human​ ​need.

It is my hypothesis that the model of the strange attractor is a provides a fruitful way to
characterize the relation between law and spontaneity in Peirce's metaphysics. Conceiving of
the system of physical laws as describing a strange attractor in phase space does justice to and
is able to explain the combination of spontaneity and law as two real manifestations of the
structure​ ​of​ ​the​ ​cosmos.​ ​It​ ​is​ ​a

conception in which law is perdurable, and thus able to answer the demands of inquiry, while
at the same time being adaptive and open to the initiatives of growth given by events of
Spontaneity. This conception may be compared to the idea of a city, something which has a
recognizable "shape" and identity but which nonetheless changes and grows. Alexander
Argyros​ ​describes​ ​this​ ​well​ ​in​ ​his​ ​book​​ ​A​ ​Blessed​ ​Rage​ ​for​ ​Order,

Clearly, a city is a dynamical entity, interacting constantly with other cities or


countries and with its own microstructure. In fact, a city is a system both
constituted and supported by feedback from the outside and the inside. A city is a
vague, amorphous, and flexible system (where does New York City end?), yet it is
undeniably real (as a New Yorker, I can attest to the overwhelming reality of New

130
York). Although part of many vertical and horizontal networks and ever in
dynamic flux, a city tends to have a real and unmistakable identity (few people
confuse New York with Paris). And finally, some cities, such as the world's great
cities, have greater depth, and a concomitant increase in identity, than lesser cities.
This depth, stemming from their history and from the layers of human work,
creativity, and imagination that combine to perform the city into being, is palpable
and goes a long way toward explaining why some cities are more interesting,
seductive, and beautiful than others. (Like cities, concepts or ideas are fractally
self-similar, dynamical systems whose shape or amorphousness is the result of a
dialectical feedback synthesis with both their environment and their history. Like
cities, ideas maintain a fluid identity that is as inimical to Derridean dissemination
as it is to metaphysical closure. Like cities, ideas are perdurable but probably
never permanent. Like cities, ideas are nodes in communication links with other
ideas, other people, and sociotemporal entities. And like cities, ideas can perhaps
be best conceptualized as attractors, or eddies, in the turbulence of personal, social,
historical,​ ​and​ ​evolutionary​ ​processes.87

As such, the world is neither wholly determinate nor arbitrary, but an irreducible combination
of​ ​the​ ​two,​ ​as​ ​the​ ​fluid​ ​yet​ ​identifiable​ ​character​ ​of​ ​the​ ​city​ ​above​ ​illustrates.

Before I move to conclude my thesis I want to consider the other sense in which I
couched my discussion of chaos theory. Thus far the discussion as been comparative and
illustrative in character but I want to consider further the extent to which chaos theory
provides confirmation of Peirce's speculations. As I have mentioned, philosophy moves with
a certain priority. The logical and categorial analyses which guide the investigations of the
special sciences cannot themselves be grounded by the use of those sciences. In asking this
question of chaos theory, though, I propose nothing so egregious as the foundation of our
philosophy, but only the inductive confirmation of a certain part of it. First, my proposal does
not "make or break" Peirce's speculations. His conception of metaphysics is based on the fruit
of his logical and categorial analyses, and though it remains hypothetical, its foundation is
solid enough to warrant its consideration as a guide in inquiry, with or without chaos theory.
Secondly, what I am proposing about chaos theory in this context is given not as a vital link
in the chain of knowledge, but as one of the many slender strands of fiber, as Peirce said, that
make​ ​it​ ​up.

As I have mentioned, Peirce himself conceived of his metaphysics as admitting of


empirical confirmation, a fact which seems only natural. If a metaphysics says that the

87
​ ​Alexander​ ​J.​ ​Argyros,​ ​A​ ​Blessed​ ​Rage​ ​for​ ​Order​,​ ​(Ann​ ​Arbor:​ ​The​ ​University​ ​of​ ​Michigan​ ​Press,​ ​1991),​ ​p.​ ​300.

131
universe has such and such a character then it is logical that investigation would come across
instances that either confirm or deny it. This is far from taking the results of science and
making our philosophy conform to it. Recalling the words of Gallie then, that given the
continued progress of scientific investigation future students of philosophy may be better
situated to appreciate the elements of Peirce's cosmology, I believe that chaos theory, and
specifically the notion of the strange attractor, provides inquiry with a substantive basis for
considering the plausibility of Peirce's metaphysics. A fundamental theme of Peirce's
philosophy derives from his critique of Cartesianism, that, as opposed to an inside/outside,
mind/matter dualism, the universe is of a continuous whole, differentiable in terms of
categories which define, not in terms of substance, but in degrees of relation. As descriptive
of all phenomena the categories necessarily predict a universe which displays possibility
(chance), facticity (existence), and regularity (law) in a combination reducible to neither one
or two of them alone. This is precisely the kind of universe that investigations in chaos theory
have discovered. The continuous holistic quality of the cosmos is manifested in the sensitive
dependence on initial conditions that chaotic systems display. A system cannot be understood
in complete isolation from those that surround and interact with it. Additionally, the order
that chaos theorists have found in chaotic systems using strange attractors models strikingly
the kind of relationship between spontaneity and law that Peirce theorized. As Peirce said, his
hypothesis "makes room for a principle of the tendency to form habits, or for a kind of
spontaneity that is to some degree regular" (6.63). In his cosmology chance and order are
reconciled in a manner which makes the discovery of swirling fractals and strange attractors
not​ ​so​ ​strange​ ​after​ ​all.

132
Conclusion

In​ ​a​ ​1904​ ​letter​ ​to​ ​William​ ​James​ ​Peirce​ ​wrote,

I want to thank you for your kind reference to me in your piece about Schiller's
Humanism. . . . The humanistic element of pragmatism is very true and important
and impressive; but I do not think that the doctrine can be proved in that way. The
present generation likes to skip proofs. I am tempted to write a little book of 150
pages about pragmatism, just outlining my views on the matter, and appending to
it some of my old pieces with critical notes. You and Schiller carry pragmatism
too​ ​far​ ​for​ ​me​ ​(8.258).

Peirce complained that James' popularized version of pragmatism had been elevated to a
self-sufficient metaphysical principle, cast off from the broad moorings of his philosophy as a
whole. Understood as such the maxim led to the view of action as being the end of man, a
conception commensurate with the nominalism that he despised. The pragmatic maxim plays
such a crucial role in Peirce's philosophy because it helps mediate the critical interface
between inquiry and the world, between signs and interpretants. As we saw, Peirce conceived
of the tendency toward generalization, the growth and association of ideas into what he called
"concrete reasonableness," as the defining character of the movement of the universe through
time. As a continuous part of the universe we partake in that growth to the extent that we
model the laws that govern nature in our own habits of inference. How this is achieved is the
fundamental question guiding Peirce's thought. This end is brought about in the process of
inquiry, specifically in the locus of abductive inference, for Peirce the only logical procedure
which introduces anything new. Achieving control over such inferences then is critical to the
growth of knowledge and of Thirdness in general. It was as a tool in achieving this control
that Peirce offered the pragmatic maxim, as a means for determining the admissibility of
hypotheses.

Because the maxim directs our attention toward the practical consequences of a
conception it is understandable why James and others reified action as the be-all and end-all
of man. It is understandable because, semiotically, it stands as a possible interpretant to the
maxim as a sign. But this interpretant will itself act as a sign, creating a constellation of ideas
and associations of which inquiry will demand explanatory power. On Peirce's reading
pragmatism so understood goes only so far in making explicable our experience, just as the
notion of determinism does in physics. Inquiry will ultimately be thwarted by understanding

133
pragmatism this way because in the course of achieving its aims it makes assumptions about
the real which an inadequate conception of the maxim bars from the outset. The real is not a
mere collection of "Seconds" but is of the nature of a continuous generality. For Peirce then,
the meaning of a concept cannot be reduced merely to action, but rather to habit, to
something which nothing empirical can correspond. Regarding the meaning of a concept, it is
"the deliberately formed, self-analyzing habit . . . [which] is the living definition, the veritable
and final logical interpretant." To give meat to this conception however requires the
contextualization of it in the normative sciences and the metaphysics of evolution. The latter
explicates the nature of the real implied in inquiry's aim, and the former defines the
normative​ ​rules​ ​by​ ​which​ ​inquiry​ ​can​ ​achieve​ ​that​ ​aim.

The question of my thesis regards the success with which Peirce was able to reorient the
broader conception of his pragmatism in the context of his later thought. In his letter to James
he implied that a proof of pragmatism was possible, though not by James' conception. If
Peirce offered a proof, it has to be one conceived on the large scale of his philosophy as a
whole, and one loosely defined, for given his doctrine of Fallibilism, a rigorous Cartesian-like
proof is not possible. Rather, I view his efforts as contributing to the "cable of knowledge" he
spoke of, "whose fibers may be ever so slender, provided they are sufficiently numerous and
intimately connected" (5.265). Peirce did not completely transform the pragmatic maxim. He
still enjoined us to clarify the meaning of concepts by deriving conditional expectations that
would be entailed in their belief, but then to realize that these practical consequences, that
action, is not sui generis but presupposes goals or ends. The discussion of ethics and
aesthetics brings to the consideration of hypotheses what Peirce called a fourth grade of
clearness, directing attention to the manner in which inquiry's aims are realized, that is, to the
growth​ ​of​ ​generality​ ​in​ ​the​ ​formation​ ​of​ ​habits​ ​of​ ​inference.

Can one say that Peirce was successful in this? By what measure can it be said that
James' version of pragmatism is wrong and Peirce's right? Clearly, at least in the context of
Peirce's philosophy, no such determination can be made. We live in a sign-interpreting
community of inquirers whose inferences are infallible only at the infinitely remote
convergence of opinion in the future, a point never to be reached. Of course this is certainly
not to say that inquiry stumbles along in the dark. Fallibility is only epistemological, for as he
said, there are a very great many things on which consensus has already been reached. Its just

134
that Peirce's semiotic transformation of Kant makes offering a proof for that simply out of the
question, for there are no transcendental faculties whose deduction could offer that. So what
criteria are to be used in assessing Peirce's work? It is clear that his theory, that any theory,
must have explanatory power. The question is, does Peirce's reorientation of the pragmatic
maxim​ ​make​ ​more​ ​fully​ ​sensible​ ​our​ ​experience​ ​as​ ​inquirers?

My judgment is that it does. There may be disagreement over certain of his


metaphysical speculations, or re-evalutions concerning the role of the normative sciences, but
that he discerned the need to explicitly address these matters at all belies the far-reaching
effect of his logical analyses. At the far ends of Peirce's architectonic lie the highly abstract
studies of mathematics and phenomenology, on one end, and his speculations on
metaphysical cosmology, on the other. In the center, around which almost all strands of his
thought can be said to ultimately revolve, is the logic of inquiry, the liminal site of contact
between us and the world we know. The achievement of his pragmaticism is to have
reoriented the referent of inquiry from things and events, to the habits and regularities which
govern them. Discern the practical consequences of a belief he says, but don't stop there, for
these are only markers, tools that aid in the work of thought. When clarifying concepts we
should resist the nominalist urge to conflate meaning with utility. The consequences which
describe its practical use help us to distinguish it from other concepts, but by paying
sufficient heed to logic we will see that it is the disposition to act, a habit, which is definable
only in terms of a "would-be" and which no number of events could exhaust, that is the real
referent​ ​of​ ​meaning.​ ​As​ ​I​ ​quoted​ ​Peirce​ ​earlier,

Pragmaticism makes thinking to consist in the living inferential metaboly of


symbols whose purport lies in conditional general resolutions to act. As for the
ultimate purpose of thought . . . it is beyond human comprehension; but according
to the stage of approach which my thought has made to it . . . it is by the indefinite
replication of self-control upon self-control that the vir is begotten, and by action,
through thought, he grows an aesthetic ideal, not for the behoof of his own poor
noddle merely, but as the share which God permits him to have in the work of
creation​ ​(5.402).

Peirce uses the word 'metaboly' to describe thinking, a word connoting the dynamic,
catalyzing activity of sign-use. It is appropriate because 'metabolism' describes those
processes necessary for the maintenance and flourishing of life, and for Peirce thinking is just

135
that, something vital and growing. Employing the pragmatic maxim without consideration for
the conditions which provide for this growth makes of the process of inquiry a stunted
enterprise. Clearly, signs will continue to develop and become more habitually ensconced in
our patterns of inquiry, but without the heuristic guides of the normative sciences and
metaphysics, the real which inquiry investigates will continue to elude us lest unwittingly we
chance upon it. Given an infinite time this near generalization of habit would indeed come
about, but Peirce is concerned with the economy and efficiency of thought. It is by the
"indefinite replication of self-control upon self-control," he says, "that the vir is begotten,"
that is, to speak broadly, that inquiry achieves its aims. This self-control certainly involves the
use of the maxim, but, as he says, "not for the behoof of his own poor noddle merely." That
which is begotten is not merely the man himself, limited by his horizons, but the work of
creation itself, the community of inquirers understood as encompassing not just humans but
all​ ​strata​ ​of​ ​the​ ​universe​ ​insofar​ ​as​ ​it​ ​is​ ​a​ ​dynamically​ ​evolving​ ​and​ ​growing​ ​system​ ​of​ ​signs.

Peirce's later thought has been criticized because it forsook the instrumental limits
imposed by the pragmatic maxim and gave scope to pragmatically unjustifiable speculation.
To conclude my assessment of his pragmaticism I offer two responses to this charge. First,
Gallie was right that future students of philosophy may be better able to evaluate Peirce's
insights. Should we assess Peirce's metaphysics with the pragmatic maxim and ask if it is an
hypothesis admissible for the work of inquiry? I think the answer is yes. I believe my
discussion of chaos theory provides the kind of "conceivable effects" that the maxim directs
us to consider. To recall, the maxim is not some magical formula for the discrimination of
truth, but merely a tool in the facilitation of inquiry. In light of what chaos theory strongly
suggests about the dynamics of the universe and the striking affinities it has with Peirce's
conception,​ ​it​ ​seems​ ​to​ ​me​ ​that​ ​as​ ​an​ ​hypothesis​ ​it​ ​merits​ ​serious​ ​consideration.

Secondly, as I have said, chaos theory is not a lynch pin in my argument in support of
Peirce, but serves only to bolster what I believe to be the already tautly woven cable of his
philosophy. Taken alone, the pragmatic maxim is a wonderfully tantalizing concept because
it neatly dispenses of so much theoretical dead wood. It was and continues to be so lauded
because it is a critically practical concept. Like a knife or a hammer, its employment yields
clear and unambiguous results. It imposes limits and defines parameters, and thus makes
apparently clear what before seemed muddled. The alluring success of contemporary

136
deconstructive practices is a good example of this and models the popularization of
pragmatism that so vexed Peirce. As purely critical, pragmatism and deconstruction can be no
more than negative in character. But Peirce believed that reflective, self-controlled inquiry
required more if it was to function in a way adequate to the fact of our ability to know the
world. Again, as he said, "after all pragmatism solves no real problem. It only shows that
supposed problems are not real problems. . . . The effect of pragmatism here is simply to
open our minds to receiving any evidence, not to furnish new evidence" (8.259). The maxim
clears away the dead wood, thus aiding in the control of inquiry, but if we are to explain how
inquiry achieves its aims, how it is that abductive insights furnish new evidence, how habits
of inference evolve and become more regular, in short, how we positively come to know the
real, we must have recourse to the constellation of concepts Peirce conceived in the general
terms of his pragmaticism. The normative sciences and the evolutionary metaphysics may
seem at distant remove from the neat precision of the pragmatic maxim, but as Peirce
understood, they are indispensable elements in any adequate explanation of the process of
inquiry.

It can perhaps be said that two concepts lie at the heart of Peirce's thought: semiosis and
inquiry. As only a triadic structure fully accounts for the interaction and development of
signs, so pragmaticism seems solely capable of explaining the demands of inquiry. In the
pragmatistic conception, inquiry is not reducible to an autonomous, logically precindable
component of experience. Rather it is an aspect or manifestation of an intricately woven
whole. The weave by which it is threaded can be said to be the categories. They pattern and
inform every aspect of Peirce's philosophy, making of it a continuous and unified whole. In
this way Peirce's account of inquiry and his speculations on metaphysics reveal a
fundamentally was ranted affinity and not the doddering wishes of a lonely secluded man. It
was only by broadening the compass of the maxim, by remaining logically consistent with
the broad outlines of his philosophy, that Peirce conceived the demands of inquiry being
explained,​ ​and​ ​of​ ​its​ ​aims​ ​being​ ​achieved.

In the end, beyond the terms of philosophical debate about pragmatism, Peirce's inquiry
itself can be seen as a sterling embodiment of the ideals he conceived of it terms of the
community of inquirers - the disinterested pursuit of truth beyond all fancies and ideological
borders. As he extravagantly said, "He who would not sacrifice his own soul to save the

137
world, is, as it seems to me, illogical in all his inferences, collectively" (2.654). In this pithy
remark we find wedded the notions of inquiry and the world in a way transcending the
short-sighted utility of a nominalist pragmatism. In the very real sacrifice Peirce made of his
life, his actual thinking exemplifies the normative self-control by which knower and known
are unified, by which the ends of thought are realized. What are these ends, he asked. We
cannot ultimately know, but using Peirce as an example we surely know that it is "not for the
behoof of our own poor noddle merely, but as the share which God permits us to have in the
work​ ​of​ ​creation."

138
Bibliography

● Almeder, R. 1980. The Philosophy of Charles S. Peirce: a Critical Introduction.


Oxford:​ ​Blackwell.
● Apel, Karl-Otto. 1981. Charles S. Peirce: From Pragmatism to Pragmaticism.
Amherst:​ ​University​ ​of​ ​Massachusetts​ ​Press.
● Argyros, Alexander J. 1991. ​A Blessed Rage for Order. Ann Arbor: The University of
Michigan​ ​Press.
● Baynes, Kenneth, et. al. 1987. After Philosophy: End or Transformation? Cambridge:
The​ ​MIT​ ​Press.
● Briggs, John, & Peat, F. David. 1989. Turbulent Mirror: ​An Illustrated Guide to
Chaos​ ​Theory​ ​and​ ​the​ ​Science​ ​of​ ​Wholeness.​ ​New​ ​York:​ ​Harper​ ​&​ ​Row,​ ​Publishers.
● Christensen, C.B. 1994. "Peirce's Transformation of Kant." ​Review of Metaphysics 48
(Sept.​ ​1994):​ ​91-120.
● Corrington, Robert S. 1993. ​An Introduction to C.S. Peirce: Philosopher, Semiotician,
and​ ​Ecstatic​ ​Naturalist.​ ​Lanham:​ ​Rowman​ ​&​ ​Littlefield​ ​Publishers,​ ​Inc.
● Davies,​ ​Paul.​ ​1988.​ ​The​ ​Cosmic​ ​Blueprint.​ ​New​ ​York:​ ​Simon​ ​and​ ​Schuster.
● Dewey, John. 1963. "The Development of American Pragmatism" ​Philosophy and
Civilization.​ ​New​ ​York:​ ​Capricorn​ ​Books.
● Ditto,​ ​William​ ​L.​ ​&​ ​Pecora,​ ​Louis​ ​M.​ ​1993.​ ​"Mastering​ ​Chaos"​ ​Scientific​ ​American.
● Earman,​ ​J.​ ​1986.​ ​A​ ​Primer​ ​on​ ​Determinism.​ ​Dordrecht:​ ​D.​ ​Reidel.
● Edie, James M. 1987. William James and Phenomenology. ​Bloomington: Indiana
University​ ​Press.
● Feibleman, James K. 1970. ​An Introduction to the Philosophy of Charles S. Peirce.
Cambridge:​ ​MIT​ ​Press.
● Gallie,​ ​W.​ ​B.​ ​1966.​ ​Peirce​ ​and​ ​Pragmatism.​ ​New​ ​York:​ ​Dover​ ​Publications,​ ​Inc.
● Gleick,​ ​J.​ ​1987.​ ​Chaos:​ ​Making​ ​a​ ​New​ ​Science.​ ​New​ ​York:​ ​Viking​ ​Penguin.
● Guyer, Paul, ed. 1992. ​The Cambridge Companion to Kant. Cambridge: Cambridge
University​ ​Press.
● Hausman, Carl R. 1993. ​Charles S. Peirce's Evolutionary Philosophy. ​Cambridge:
Cambridge​ ​University​ ​Press.
● Hayles, N. Katherine. 1990. Chaos Bound: Orderly Disorder in Contemporary

139
Literature​ ​and​ ​Science.​ ​Ithaca:​ ​Cornell​ ​University​ ​Press.

−−−. ​1991. Chaos and Order: Complex Dynamics in Literature and Science. ​Chicago:
University​ ​of​ ​Chicago​ ​Press.

● Hookway,​ ​Christopher.​ ​1985.​​ ​Peirce.​ ​London:​ ​Routledge​ ​&​ ​Kegan​ ​Paul​ ​plc.
● Kellert, Stephen H. 1993. In the Wake of Chaos. Chicago: University of Chicago
Press.
● Ketner, Kenneth Laine, ed. 1995. ​Peirce and Contemporary Thought. New York:
Fordham​ ​University​ ​Press.
● Lorenz, E. 1963. "Deterministic Nonperiodic Flow." Journal of the ​Atmospheric
Sciences​ ​20:​ ​130-41.
● Ochs, Peter. 1993. "Charles Sanders Peirce." In ​Founders of Constructive Postmodern
Philosophy: Peirce, James, Bergson, Whitehead, and Hartshorne. Ed. David Ray
Griffin,​ ​et.​ ​al.​ ​New​ ​York:​ ​State​ ​University​ ​of​ ​New​ ​York​ ​Press.
● Peirce, Charles Sanders. 1931-1935. ​The Collected Papers of Charles S. Peirce. ​8
vols. Ed. C. Hartshorne, P. Weiss, and A. Burks. Cambridge: Harvard University
Press.

----- . ​1982-1986. Writings of Charles Sanders Peirce: A Chronological Edition, ​Vols. I-IV.
Ed.​ ​Max​ ​Fisch.​ ​Bloomington:​ ​Indiana​ ​University​ ​Press.

----. 1977. The Correspondence between Charles S. Peirce and Victoria Lady Welby​. Ed.
Charles​ ​S.​ ​Hardwick.​ ​Bloomington:​ ​Indiana​ ​University​ ​Press.

----. ​1967. Annotated Catalogue of Charles Sanders Peirce. Ed. Richard S. Robin. Amherst:
University​ ​of​ ​Massachusetts​ ​Press.

----. 1955. The Philosophical Writings of Peirce. ​New York: Dover ​Publications: 1955), p.
340.

● Potter, Vincent. 1996. Peirce's Philosophical Perspectives. New York: Fordham


University​ ​Press.
● Ransdell, J. M. 1977. "Some Leading Ideas of Peirce's Semiotic", ​Semiotica, vol. 19,
pp.​ ​157-78.
● Saussure, Ferdinand de. 1960. ​Course in General Linguistics. Ed. C. Bally et. al. 5th
ed.​ ​Paris:​ ​Payot.

140
● Sheriff, John K. 1994. Charles Peirce's Guess at the Riddle: Grounds for Human
Significance.​ ​Bloomington:​ ​Indiana​ ​University​ ​Press.
● Wiener, P. P. 1949. ​Evolution and the Founders of Pragmatism. ​Philadelphia:
University​ ​of​ ​Pennsylvania​ ​Press,​ ​1972.

Additional​ ​Sources​ ​of​ ​Information​ ​On​ ​Peirce​ ​and​ ​Chaos​ ​Theory

On​ ​Peirce:

● Boler, J. 1980. Charles Peirce and Scholastic Realism. ​Seattle: University of


Washington​ ​Press.
● Bucjler,​ ​J.​ ​1939.​ ​Charles​ ​Peirce's​ ​Empiricism.​ ​New​ ​York:​ ​Harcourt,​ ​Brace​ ​&​ ​World.
● Christensen, C. B. 1994. "Peirce's Transformation of Kant", Review of Metaphysics​,
vol.​ ​48,​ ​pp.​ ​91-120.
● Colapietro, Vincent M. 1989. ​Peirce's Approach to the Self: A Semiotic Perspective
on​ ​Human​ ​Subjectivity.​ ​Albany:​ ​SUNY​ ​Press.
● Deledalle, Gerard. 1990. Charles S. Peirce: An Intellectual Biography. ​Trans. Susan
Petrilli.​ ​Amsterdam:​ ​John​ ​Benjamins.
● Descartes, R. 1985. Rules for the Direction of the Mind, in The Philosophical
Writings of Descartes. Trans. John Cottingham, Robert Stoothoff, and Dugald
Murdoch.​ ​Cambridge:​ ​Cambridge​ ​University​ ​Press.
● Esposito, Joseph L. 1980. ​Evolutionary Metaphysics: The Development of Peirce's
Theory​ ​of​ ​Categories.​ ​Athens:​ ​Ohio​ ​University​ ​Press.
● Fitzgerald, John J. 1966. Peirce's Theory of Signs as Foundation for Pragmatism. The
Hague:​ ​Mouton.
● Frege, Gottlob. 1964. The Basic Laws of Arithmetic: Exposition of the System. ​Trans.
and​ ​ed.​ ​Montgomery​ ​Furth.​ ​Berkley:​ ​University​ ​of​ ​California​ ​Press.

----. 1984. "Function and Concept", Collected Papers on Mathematics, Logic, and
Philosophy.​ ​Trans.​ ​Max​ ​Black,​ ​et.​ ​al.,​ ​Ed.​ ​Brian​ ​McGuinness.​ ​Oxford:​ ​Basil​ ​Blackwell.

● Goudge, T. A. 1950. ​The Thought of C. S. Peirce. Toronto: University ​of Toronto


Press.

141
● Greenlee,​ ​D.​ ​1973.​ ​Peirce's​ ​Concept​ ​of​ ​Sign.​ ​The​ ​Hague:​ ​Mouton.
● Murphey, Murray. 1961. The Development of Peirce's Philosophy. ​Cambridge, Mass.:
Harvard​ ​University​ ​Press.

----. 1968. "Kant's Children: the Cambridge Pragmatists", ​Transactions of the Charles S.
Peirce​ ​Society,​ ​vol.​ ​4,​ ​pp.​ ​3-18.

● Potter, Vincent. 1967​. Charles S. Peirce: On Norms and Ideals. ​Amherst: University
of​ ​Massachusetts​ ​Press.
● Putnam, Hillary. 1982b. "Peirce the Logician", ​Historia Mathematica​, vol. 9, pp.
290-301.
● Reilly, Francis E. 1970. ​Charles Perice's Theory of Scientific Method. ​New York:
Fordham​ ​University​ ​Press.
● Rescher, N. 1978. Peirce's Philosophy of Science. ​Notre Dame: ​University of Notre
Dame​ ​Press.
● Sharpe, R. A. 1970. "Induction, Abduction and the Evolution of Science",
Transactions​ ​of​ ​the​ ​Charles​ ​S.​ ​Peirce​ ​Society,​ ​vol.​ ​6,​ ​pp.​ ​12-33
● Thompson, M. 1953. The Pragmatic Philosophy of Charles S. Peirce. ​Chicago:
University​ ​of​ ​Chicago​ ​Press.

----.​ ​1978.​ ​"Peirce's​ ​Verificationist​ ​Realism",​ ​Review​ ​of​ ​Metaphysics,​ ​vol.​ ​32,​ ​pp.​ ​74-98.

On​ ​Chaos​ ​Theory:

● Bachelard, G. 1984. ​The New Scientific Spirit. Trans. A. Goldhammer. Boston:


Beacon​ ​Press.
● Barnsley, M., and S. Demko, eds. 1986. Chaotic Dynamics and Fractals. Orlando:
Academic​ ​Press.
● Bergé, P., Y. Pomeau, and C. Vidal. 1984. ​Order Within Chaos. Trans. L. Tuckerman.
Paris:​ ​J.​ ​Wiley​ ​&​ ​Sons.
● Burtt, E. A. (1952) 1980. ​The Metaphysical Foundations of Modern Science. Atlantic
Highlands,​ ​N.J.:​ ​Humanities​ ​Press.
● Conrad,​ ​J.​ ​(1914)​ ​1984.​ ​Chance:​ ​A​ ​Tale​ ​in​ ​Two​ ​Parts.​ ​London:​ ​Hogarth​ ​Press.

142
● Crutchfield, J., J. D. Farmer, N. Packard, and R. Shaw, 1986. "Chaos", Scientific
American​ ​225​ ​(December):​ ​46-57.
● Devaney, R. 1986. An Introduction to Chaotic Dynamical Systems, ​Menlo Park:
Benjamin/Cummings​ ​Publishing​ ​Co.
● Dyke, C. 1990. "Strange Attraction, Curious Liason: Clio Meets Chaos."
Philosophical​ ​Forum​ ​21:369-92.
● Fischer, P., and W. R. Smith. 1985. ​Chaos, Dynamics, and Fractals. ​New York:
Marcel​ ​Dekker.
● Ford,​ ​J.​ ​1983.​ ​"How​ ​Random​ ​Is​ ​a​ ​Coin​ ​Toss?"​​ ​Physics​ ​Today​ ​36:​ ​40-47.

----. 1986. "Chaos: Solving the Unsolvable, Predicting the Unpredictable!" In M. Barnsley
and​ ​S.​ ​Demko,​ ​eds.,​ ​Chaos​ ​Dynamics​ ​and​ ​Fractals​,​ ​1-52.​ ​Orlando:​ ​Academic​ ​Press.

● Glass, L., and M. C. Mackey. 1988. ​From Clocks to Chaos. Princeton: Princeton
University​ ​Press.
● Grassberger, P., and I. Procaccia. 1983. "Characterization of Strange Attractors."
Physicial​ ​Review​ ​Letters​ ​50:346-49.
● Haken,​ ​H.,​ ​ed.​ ​1981.​ ​Chaos​ ​and​ ​Order​ ​in​ ​Nature.​ ​Berlin:​ ​Springer​ ​Verlag.
● Hobbs, J. 1991. "Chaos and Indeterminism." ​Canadian Journal of Philosophy 21:
141-64.
● Hofstadter, D. R. 1981. "Strange Attractors: Mathematical Patterns Delicately Poised
between​ ​order​ ​and​ ​Chaos."​ ​Scientific​ ​American​ ​245​ ​(November):​ ​22-43.
● Holden,​ ​A.​ ​V.,​ ​ed.​ ​1986.​​ ​Chaos.​ ​Princeton,​ ​N.J.:​ ​Princeton​ ​University​ ​Press.
● Hunt,​ ​G.​ ​M.​ ​K.​ ​1987.​ ​"Determinism,​ ​Predictabiltiy​ ​and​ ​Chaos."​ ​Analysis​ ​47:​ ​129-33.
● Kim, Y. S., and W. W. Zachary, eds. 1987. ​The Physcis of Phase Space. Berlin:
Springer-Verlag.
● Livi, R., S. Ruffo, S. Ciliberto, and M. Buiatti, eds. 1988. ​Chaos and Complexity.
Singapore:​ ​World​ ​Scientific.
● Monod,​ ​J.​ ​1971.​ ​Chance​ ​and​ ​Necessity.​ ​New​ ​York:​ ​Alfred​ ​a.​ ​Knopf.
● Prigogine,​ ​I.,​ ​and​ ​I.​ ​Stengers.​ ​1984.​ ​Order​ ​out​ ​of​ ​Chaos​.​ ​New​ ​York:​ ​Bantam​ ​Books.
● Salmon, W. 1984. ​Scientific Explanation and the Causal Structure of the World.
Princeton:​ ​Princeton​ ​University​ ​Press.
● Skarda, C. A., and W. J. Freeman. 1987. "How Brains Make Chaos in order to Make

143
Sense​ ​of​ ​the​ ​World."​ ​Behavioral​ ​and​ ​Brain​ ​Sciences​ ​10:161-95.
● Stone, M. 1989. "Chaos, Prediction, and Laplacean Determinism." ​American
Philosophical​ ​Quarterly​ ​26:123-31.

144

Vous aimerez peut-être aussi