Вы находитесь на странице: 1из 331

Investigations in

Philosophy of Space
Elisabeth Str6ker
translated by
Algis Mickunas
~ Ohio University Press
1 Athens, Ohio London
Translation Copyright 1987 by Ohio University Press
Originally published as Philosophische Untersuchungen
zum Raum. (Frankfurt Am Main: Vittorio Klostermann, 1965)
Copyright 1965.
Printed in the United States of America.
All rights reserved.
Library of Congress Cataloging-in-Publication Data
Stroker, Elisabeth.
Philosophical investigations of space.
(Series in continental thought; v. 11)
Translation of: Philosophishe Untersuchungen zum
Raum.
Bibliography: p.
Includes index.
1. Space and time. l. Title. 11. Series: Series
in continental thought ; 11.
BD632.S6513 1987 114 86-2410
ISBN 0-8214-0826-7
Table of Contents
Translator's Preface ................................................ ix
Preface ............................................................... xi
Introduction ............................................................. 1
1. The State of the Problem ................................ 1
2. The Aim of the Investigation ............................ 3
3. Preliminary Methodological Considerations ........... 7
Part One: Lived Space ................................................ 13
Section One: Contributions to the Phenomenology of
Lived Space ........................................ 14
Point of Departure and Statement of the Problem ........... 14
Chapter One: The Attuned Space ............................. 19
1. The Concept of Attuned Space ......................... 19
2. Characteristics of Attuned Space: Fullness and
Emptiness ................................................. 22
3. Place and Position in Attuned Space .................. 27
4. Nearness and Remoteness .............................. 29
5. Movement and Orientation in Attuned Space ........ 30
6. Attuned Space as Space-Time .......................... 36
7. Attuned Space and the Experiencing Subject ......... 43
Chapter Two: The Space of Action ........................... 48
1. Preliminary Remarks .................................... 48
2. Place and Regan. The Space of Action as a
Topological Manifold ................................... 52
3. The Locus of the Subject in the Space of Action ..... 57
4. Movement and Orientation. The Space of Action
as Oriented Space ........................................ 62
5. The Problem of the Way ................................ 71
6. Nearness and Remoteness in the Space of Action .... 75
7. Summary ................................................. 81
Chapter Three: The S pace of Intuition ....................... 83
1. Terminological Clarifications ........................... 83
2. The Space of Intuition as a Phenomenal Multitude
of Points ........... ~ ................................. .' ... 85
V
vi Table of Contents
3. The Lived Body as the Center of the Space of
Intuition .................................................. 89
4. The Oriented Space of Intuition ....................... 90
5. Spatial Depth and Perspectivity ........................ 93
6. The Finitude of the Space of Intuition ............... 109
7. The Other in My Space of Intuition. Questions of
Homogenization ........................................ 113
8. Open Questions ........................................ 118
Chapter Four: Modally Distinct Sensory Spaces ........... 120
1. Visual Space ............................................ 120
2. The Visual Field ........................................ 124
3. The Problem of Tactile Space ......................... 126
Section Two: Questions of Space Constitution ................ 138
Chapter One: Corporeity and Spatiality ..................... 138
1. Methodological Survey ............................ ' .... 138
2. The Lived Body and the Physical Body in their
Relationship to Space .................................. 140
3. The Lived Body and Consciousness .................. 148
Chapter Two: The Space of Movement and Objective
Space .......................................... 152
1. Spatial Structure and Corporeal Facticity ........... 152
2. The Problem of Empty Space ......................... 160
3. Concluding Observations on Lived Space ........... 169
Part Two: Mathematical Space ..................................... 173
Introductory Remarks ............................................ 174
Section One: Preliminary Phenomenological Observations . 176
Chapter One: Space as a Thematic Object of
Consciousness ................................. 176
1. The Space of Intuition as a Limit Case of Lived
Spatiality ................................................ 176
2. The Topological Structure of the Space of Objects . 179
Chapter Two: Basic Trends of Mathematization ........... 184
1. Morphological and Mathematical Determinations
of the World of Things ................................. 184
2. The Problem of Mathematical Ideation .............. 190
3. Symbolic Intuition (Pictorial Symbolism) ........... 194
4. Signitive Symbolization of Geometry ................ 200
5. The Constructive Character of Geometric
Objectivity. Geometry as a Demonstrative Science . 211
6. Summary ................................................ 221
Table of Contents vii
Section Two: Euclidean Space .................................. 225
Chapter One: Phenomenological Access to Metrics ........ 225
1. Formation and Relationship. The Primacy of
Relationships ........................................... 225
2. The Line Segment as a Fundamental Metric
Formation ............................................... 228
3. The Line Segmentas an Invariant of
"Movements" ........................................... 231
4. The Concept of Movement as a Leading Concept
of the Theory of Invariants ............................ 236
Chapter Two: Euclidean Normal Space ..................... 239
1. The Concept of Mathematical S pace (Preliminary
Conceptual Clarification) .............................. 239
2. Normal Space (Euclidean Space of the
Topological Type of the Open Plane) ................ 246
3. The Question of Intuitability in Euclidean
Geometry ................................................ 252
Chapter Three: Euclidean Spaces with Topological
Anomalies ................................................... 258
1. Extension of the Mathematical Concept of Space ... 258
2. Clifford-Klein Spaces .................................. 259
3. Clifford-Klein Spaces as Euclidean Normal Space.
Founding Relationships ............................... 260
Section Three: Non-Euclidean Spaces .......................... 265
Chapter One: Fundamental Questions of Non-Euclidean
Geometry .................................................... 265
1. The Parallel Postulate. Historical Origin and
Development ............................................ 265
2. Constitutive Problems of the Parallel Postulate ..... 269
Chapter Two: Foundational Problems of Hyperbolic
Geometry .................................................... 2 7 5
1. On the Metrics of Hyperbolic Geometry ............. 275
2. The Kleinian Model. Phenomenological Analysis
of the Model Conception .............................. 278
3. Hyperbolic Geometry and the Space of Intuition ... 281
Chapter Three: Riemann's Geometry ........................ 285
1. Riemann's Point of Departure. The Metric
Fundamental Form ..................................... 285
2. Riemannian Spaces. Brief Mathematical
Characterization ........................................ 287
3. Curvature and "Curved Spaces" ...................... 291
viii Table of Contents
4. The Question of the Existence of the Mathematical
Point ..................................................... 296
Concluding Observations ........................................... 304
Works Cited and Consulted ........................................ 309
Register . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Translator' s Preface
Professor Dr. Elisabeth Stroker is a noted philosopher of the
history of science and a leading thinker in the current development
of philosophy of technology, inclusive of ethical issues attendant
upon science and technology. I am grateful to Professor Stroker for
the trust and patience she maintained in me during the protracted
process of translation. This is more so in face of the importance of
her work in the philosophy of space. It contains the best possible
phenomenological and critical analyses of issues concerning the
"experience" and the "being" of space, and the most outstanding
and precise theoretical investigations of the problems in conceptu-
alizing space. Professor Stroker's theoretical investigations include
the development of the entire Western philosophical tradition
concerning the conceptions of space, ranging from Euclid all the way
to the geometric-mathematical problematics of Riemann. To say the
least, Professor Stroker's work is both and scientif-
ically fundamental. It sets the standard for any future investigations
in the philosophy of space.
The success of the translation must be attributed to Dr. Elizabeth
Behnke who, under the auspices of a copy editor, accomplished a
monumental task. Her indispensible cooperation, her inexhaustible
patience with every term and phrase, are completely intertwined in
the translation. Indeed, Dr. Behnke was part and parcel of making
the text intelligible toan American reader. Many thanks.
Preface
This investigation is m y habilitation work, revised for print, which
was presented to the philosophy faculty of the University of Ham-
burg in the summer semester of 1963. It was begun thanks to a
generous stipend from the German Research Society. Its preparation
was enhanced by professor Dr. Wolfgang Wieland of Hamburg
University. 1 am grateful for his interest and valuable suggestions. 1
also owe my gratitude to Professor Dr. Carl Friedrich Feiherr van
Weizsacker and, not the least, to Professor Dr. Gnther Patzig for
their enhancement of the work through their encouragement and
critical remarks.
The first conceptual inception of this writing reaches back toward
the years of my study at the University of Bonn. The lectures and
works of Oskar Becker impressed me by their masterful subtleties of
philosophical analyses of problems and pointed me, for the first time,
toward phenomenological questioning. Less obvious is the imput in
this work that was acquired from Theodor Litt. He has impacted and
swayed my first philosophical studies; he played a vital role in the
proposed investigation. While the path followed in this work re-
mained distant to him, this distance allowed him to survey the work
more intensely, allowing him unrestrained and open criticism.
Hence the will of a teacher to demand, in which the power of his
thought was most vital, became a most conspicuous part of this
work. His frequently expressed wish to see the work attain its
proposed aim, was not granted to him. As his last student, 1 dedcate
this work to his memory.
Hamburg, September 1964 Elisabeth Striker
Introduction
1. The State of the Problem
The philosophical treatment of space faces a problem whose
tradition is as old as the entire history of philosophy. As an object of
metaphysical and naturalistic philosophical speculation since the
pre-Socratics, space seems to offer an almost inexhaustible variety of
aspects throughout the historical changes of cosmogonies, systems
and conceptual positions. What is given in them to date as a
"theory" of space turns out to be a sublimated sediment of histori-
cally acquired concepts from whose entire intellectual content space
assu'mes its decisive outlines. This is true for the finite central-
peripheral spatial cosmos of antiquity and its modifications in high
scholasticism, as well as for the infinite, homogeneous space of
Renaissance philosophy; the same is true for the absolute space of
Newton no less than for the much acclaimed "renaissance" of the
Aristotelian conception of space in the field theories of modern
physics.
With the development of modern natural sciences, thinking con-
cerning space acquires new impulses from experimental research. 0\- eLt- _
Non-Euclidean geometries and their extension of the concept of
space have contributed to the discussion of the problem from the
mathematical side since the end of the eighteenth century. History
shows that from then on philosophy and the exact sciences not only
diverge, but also continuously confront one another in the contro-
versies concerning epistemic claims arising from the treatment of the
problem of space. The fact that since the beginning of the nineteenth
century this controversy has been oriented almost exclusively by
Kant's conception of space indicates, on the one hand, the extraor-
dinary fertility of the Kantian a priori and, on the other hand, has its
basis in the fact that the latter bears unmistakably within itself the
properties of the universal structure of Newtonian space whose
validity has become increasingly questionable. The most recent and
in tense encounter between philosophy and science was precipitated 1 [Aroi:J 1}-
by the theory of relativity at the' dawn of this century. It is { h1
revolutionary: it transformed the entire physical world image and )
1
rcW[\'Of\g'vv.-1['6.
<S fU U. , .. , "'f"fA .t( .
2 Introduction
renewed the demand for a reply to the philosophical question of
space.
Subsequently, it seems as though the problem of space has become
merely an occasion for a reciproca! glance across the boundaries
between philosophical and scientific research. Meanwhile, the phi-
losoiJhy of our day begins the discussion from another point of view.
With Jpward the world
prior to. gart __
things. .tbeir
their everyday--existence. Wherever existential philosophy deals
with space within the framework of it
only recognizes as philosophically significant the space of "Dasein"
---? experienced in its everydayness apart from scientific conceptions.
Heidegger's differentiation of "being in" from being "in" something
attempts to distinguish the traditional categorical conception of
space from space as existential; Merleau-Ponty's alternative concept
of a "troisieme spatialit", in contrast to physical and mathematical
spatiality, strongly suggests a domain of spatial research inaccessible

"$
)/;.(L.::
to the mathematical sciences. Perhaps the solution to the riddle of )!
space should be sought here, solution airead y clearly seen by Kant, V..?l
for in arder to eluciate the gossibility of things, he presented space
co1J.!!i1!2!1 for su eh -
Influences of existential-ontological and existential-phenomeno-
logical efforts are particularly obvious in the area of contemporary
psychology. Although the naturalistically oriented scientific psy-
chology had begun to investigate space perception experimentally
and genetically over a hundred years ago, more recent research
follows the various analyses of "space experience" under the aspect
of holistic psychology, phenomenological description, and method-
ological hermeneutics. In contemporary psychopathology there is a
notable emphasis on the abnormal experience of space, leading toa
multi-faceted analysis of "spatial disturbances." Such research is not
too encouraging, since during the last decades it has led to a
confusing multitude of various "spaces." Yet, at the same time, this
points to numerous efforts under various methodological aspects to
master the complex and multi-layered problem of space.
If the entire range of problems were finally surveyed, then imme-
diately specific questions would have to emerge. With what justifi-
cation does one speak of "spaces"? Kant was of the opinion that
space must be necessarily unitary, and that all talk of spaces made
sense only if they were parts of the same space. Has this claim lost
its validity today? The space of intuition and the space of mathemat-
Introduction 3
ics cannot be arranged next to each other and are not conceivable as
parts. If there were a more encompassing space containing them as
parts, then its structure should be accessible to investigation; if not,
then it would have to be seen as irrational, and the genuine, unitary
space would be inaccessible to thought. Or is it that the title "space"
can be genuinely claimed by only one of the regions currently called
spaces? Or perhaps in accordance with its meaning, space ought to
be sought and found apart from sciences in our direct presence to
and orientation toward the world. In this case, what is always present
in the pre-reflective, everyday consciousness, what is grasped
unthematically and understood as space befare the inception of
conceptual activity, would have to be elevated to reflective light by
philosophical investigation. Furthermore, the investigation would
have to justify to what extent any further talk of spaces is nonsen-
sical; it would have to decipher the source of error, expose the
speculative willfulness and inappropriate systematization where
thinking, too intent on a unitary vision, grasps aspects of a most
heterogeneous kind under the concept of space.
If the talk about spaces is valid, then we should ask about the
differences and commonalities of their structures. Furthermore, it
EJ.ight of does __
possess, /
:@!!.li!Y? Phosophy must not only seek an answer, it must also take //
a stand with regard to the sense of question itself.
2. The Aim of the Investigation
The ultima te aim of this investigation pertains to theory of science;
it purports to be a contribution toward a philosophical grounding of
geometry. The concept of the theory of science must nevertheless be
taken in a broader than usual sense. Since the end of the nineteenth
century, the theory of science has been understood predominantly as
methodology or as applied logic. Thus, it analyzes the modes of
operation of a science, observes its specific concept formation, and
investigates its conceptual assumptions, its modes of deduction and
the type of its laws. Ideally, each science regards itself as method-
ologically unified. That is to say it views itself as an embodiment of
statements whose interrelationship is based either on a finite, even if
numerically and typologically variable, quantity of irreducible prin-
cipies or on a limited number of distinct, although factually and
logically unifiable types of operations. The first is concerned with
axiomatic disciplines, the second with experimental sciences.
4 Introduction
Within each individual science, it is possible to investigate and
compare the structures of relatively closed partial systems of prop-
ositions, "theories," according to the type of their basic presuppo-
sitions and procedures. If the comparison of methods is extended to
different sciences, then we would devise a common methodology as
a theory for a specific group of sciences. It could function as a
methodological framework for such groups as natural or human
sciences. As a general and inclusive methodology of the sciences, it
is subsumed under the idea of the unity of all possible forms of
theory and thus under the essential explication of science in
general.
Yet a theory of science is not exhausted by pure methodology. As
a reflective effort toward the structural clarification of the sciences in
the broadest sense, it must remember that sciences are ultimately
sciences of beings, oriented toward what "is," what "exists." Indi-
vidual scientific questions asking what is, and how it is, are oriented
ontologically toward qualities, quantities, relations, and causal and
functional laws. In contrast, a theory of science requires another
position. This position is concerned with a factual area of research,
not taken in the simple objective sense of its immediate "being," but
in a mediated sense: something existent that is "on the way" to a
specific methodological exposition. Only in this modified sense
does this area become an object for the theory of science. The efforts
of a theory of science go beyond a mere methodology in yet another
respect. Method is not merely about something existent that is to be
known; it is equally a method of the interrogator who knows and
wishes to know more about what there is. Conceived in this manner,
a theory of science must take its point of departure from an entirely
different ground than from mere methodology. In contrast to meth-
odology, a theory of science assumes at the outset a more encom-
passing task in a twofold manner.
Taking science in its objective relationship to the subject matter as
well as in its relation back to specific forms of the activity of
subjectivity, a theory of science faces a problem that can be called
ontological in a specific sense of contemporary philosophical
thought. What we mean here is a problem emerging froma continu-
ing interrogation of the subject matter whose aim is a universal
explication of the "sense" of being. This process must pay constant
attention to the being of the interrogating subject (Husserl's transcen-
dental ego, Heidegger's Dasein). With regard to our present investi-
gation and its precise limitation to its thematic object, this means
that a theory of science is necessarily confronted by the questions as
Introduction 5
to what space might be and in what sense unqualified assertions of
its existence can be given.
Moreover, in contrast to methodology, a theory of science cannot
begin with the established sciences in arder to discover their
methodological structure. Rather, in its attempt to make the method
comprehensible, it must return to tl;w already given relationships
and states of affairs from which the scientific questions originate.
With specific regard to space, this means that its treatment in a
theory of science cannot begin with the abstract mathematical
spaces-assuming for the time being their plurality-but with the
spatiality given and experienced prior to all sciences. Thus this \lve-J
investigation begins with a detailed treatment of the "lived" space.
li)f(). ce._
lt is based on the view that a solution to the question of the meaning
a_QQEl.!t:lnce of the "abstract" spaces of the specjal sciences and
can be sought only on the grounds of a prior investigation
is_J;:Q:::given
an object of This corresponds to the "
late Husserl: all objectivity of science is
necessarily related to corresponding elements found in the "lived
world." This relationship founds the validity of the sciences. lt thus
remains to be explicated "how all the self-evidence of objective-
logical accomplishements, through which objective theory (thus
mathematical and natural-scientific theory) is grounded in respect to
form and content, has its hidden, source of grounding in the
ultimately accomplishing life, the life in which the self evident
givenness of the life world forever has, has attained, and attains
anew its prescientific, ontic meaning. "
1
The following work is justifiable only as an attempt to clase a gap
that has not yet been successfully filled, despite a series of valiant
attempts. lt propases to develop the problems of space under the
main notions suggested above. lt will provide a critica! illumination
and mutual evaluation of the many scattered concepts found in
diverse branches of research, without subsuming their heteroge-
neous results and diverse theoretical positions under a forced
unification.
The work does not make any claims to completeness. lt deliber-
ately limits its analyses to the problems of lived space and the spaces
of geometry, i.e., the mathematical spaces as free geometric mani-
folds. Thus it excludes an entire domain of questions concerning the
practica! use of geometric structures in physics. "Physical space,"
1. E. Husserl, Krisis, 34d; see also 38.
6 Introduction
which actually should be added, remains here only as a requirement
to be fulfilled by a subsequent and more exhaustive investigation.
Our work takes into consideration tasks whose complexities by far
surpass our efforts. At the very outset we are not certainvvhether and
The -free
ideal formations of geometry, with regard to their pure geometrical
existence, to be investigated in the second part of this work, are
extremely difficult to explicate, especially in their relationship to the
things of "real" space. This relationship is particularly difficult to
explain with regard to physical bodies and their spatial interaction
in physics. The term "application" not only does not explain this
relationship, but actually hides various problems. If geometry could
be equated with a tool of physics that is constituted and, so to speak,
prepared by the mathematician, then the problem could be resolved
quite readily. Yet the relationship between geometry and physics is
not that simple. The elucidation of such a relationship must remain
the task of a more specialized work.
Our investigation may be called systematic only to the extent that
it is not conceived as a history of problems and is therefore not to be
subsumed under the criteria pertaining to the latter sort of study.
While such a distinction is justifiable for science, we admit that it
may be dangerous and misleading in philosophy. To propase a sharp
separation between the systematic and the historical aspects of
philosophy would result in an obfuscation of the fact that each
"system" contains at least as much "history" as the conceptuality it
uses rests in historical origin and retains the heritage of a long
tradition. At the very outset, the systematic development of a
problem is not primarily determined by its historically evolved
status as a scientific problem; rather, it is already conditioned by a
situationally determined horizon of the pregiven comprehensibility
and unproblematic self evidence of everyday life. Though we may
not recognize or even admit such obvious and everyday intelligibil-
ity, it is no less historical and historically constructive.
This investigation intends to deal systematically with the problem
@ of space as it is articulated in contemporary views and made
accessible by current methods. Yet it is to be remembered that the
investigation does not merely rest "on the ground" of a tradition that
\:)(]IV\ is to be surveyed in its temporal unfolding, useful perhaps only for
dexographic interests; rather, the work is "taken up" with a tradi-
tion, to the extent that the tradition eifers- into ___ the )i()lJTms
fi1Ve-stigated, as is the case for the second part of this work, and to the
extent that the mode and manner of questioning are shaped by the
Introduction 7
tradition. After all, we raise questions about space not because we
know nothing of it, but because we always ha ve a prior acquaintance
with it. It is already grasped and laid out in sorne sense befare we
begin to ask questions about it. Its presence must be such that it is
open to our questioning. Although our effort fails to do justice to
history, it is aware of its indebtedness to history.
2
At this juncture we must briefly indicate a specific difficulty. The
plan of our investigation must be limited to space without due
consideration to time. At first glance, leaving out time seems to avoid
an unnecessary complication; yet a closer look reveals a genuine
lacuna. S_llilce and time constitute a unified whole and thus are
1
related essentially. Such-a relatl:msiifp-does--not appear siiiPTY
because both are "forms of intuition." Their relationship does not
become obvious by a simple declaration that space is "the arder of
one next to the other" and time is "the arder of one after the other."
This usual conception of SJ2!!Ce and
The unity does not lie in the simple fact that each case of one next
to the other includes a specific "temporal point." The language and
the aims of modern physics, expressed in Minkowski's space-time
continuum, has done proper justice to this unity. Yet the intertwin- ,()
ing of space and time is more intimate and fundamental in the
domain of daily life. Without going into greater depth, the present
work can only indicate the crucial domains of such an interrelation-
ship.
3. Preliminary Methodological Considerations
An investigation whose intent is to unfold the problem of space in
its entirety requires, first of all, methodological guarantees.
Though our endeavour will subsequently be delimited more
precisely, it can initially be characterized as phenomenological.
Today phenomenology flourishes in a multitude of trends, and no
"- - .-" -....._" ... -r--" ---
2. For the historical treatment of the problem of space, we can mention
the works of W. Gent, H. Conrad-Martius, K. Deichmann, H. Heimsoeth, A.
Koyr, and M. particular the latter presents an excellent
treatment of the physicalistic history of the problem, ranging from the
natural philosophy of the Greeks to the general theory of relativity. The work
of Koyr is limited to the historical period between Cusanus and Leibniz,
although it is most trustworthy for this historical period. The intent of E. y,))
Fink's worMs not primarily "historic-doxographic"; rather, it takes up the ---
entire ontd'(ogical tradition und_E)!'_ t_hg le_ading aspe_<_;t1; time, and
--- -
hf'k
8 Introduction
longer admits of a unitary procedure. Nowhere has it taken over the He_..,
1
conception of its celebrated past master Hegel. For Hegel, phenom- J
enology was a way of "apparent knowledge," comprising a prelim-
inary step with which philosophy should begin in order to attain
true knowledge as the end phase of a dialectical movement. But
Husserl did not follow in Hegel's steps. Since Husserl, phenomeno- Hv\\
1
logical philosophy considers its task to be "pure description" of a
"given." The given ranges from apure act phenomenon,. within the
framework of the Husserlian reductions, to the "being-in-itself"
within the unreflective attitude of naive realism. Even existential
ontology is concerned with seeing entities as they are "in them-
selves." The efforts of existential ontology are centered on "allowing
that which is to come to manifestation." In its "hermeneutic" }-).., l,
procedure it adheres consciously to the Heideggerian formulation of
"the tautology of a descriptive phenomenology."
3
To elucidate the value of phenomenological procedure for its own
sake would be like carrying coals to Newcastle. Certainly it has
established a method that is responsible for subtle analyses and far
reaching insights in all areas of contemporary philosophy. Yet this
; raises a serious question: can philosophy continue to use this
procedure without running the risk of eliminating the possibility of
self-critique as the fundamental element of its life? Philosophy
requires self-critique more than does science, for in contrast to many
of the individual sciences, philosophy does not have a constant
corrective for its statements in experience. Philosophy can realize
such a critique only if it elevates reflective thinking to its ultimate
possibilities. Thus it must continuously submit the capacity and
limits of its own method to the light of critical self-reflection. Even
if at the first glance the given seems indubitable and self-evident,
philosophy must admit at the very outset that its "given" is
completely conditioned by its formulation and thus exposed to
specific limitations, which are justifiable only by the choice of a
point of departure; the given can only be challenged at the point of
departure. Thus the Husserlian "principie of all principies" for
phenomenology, namely that "every originary presentive intuition
is a legitimizing source of that "everything originarily
... offered to us in 'intuition' is to be accepted simply as what it is
presented as being, but also only within the limits in which it is
presented,"4 must not only be maintained most precisely while
3. M. Heidegger, 7c.
4. E. Husserl, Ideen I, 24.
Introduction 9
working with specific descriptive analyses, but must be respected
more precisely than was done by Husserl himself in the process of
his own phenomenological investigations. Moreover, a
lil!l:it1l .. ()f the same time the
limits of phenomenology itself? Since by its owitenets it is.orierite(f
''the thingstliemselves'r, and devotes itself entirely to indi-
vidual researches in accordance with the prescript of a science of
science, it is caught in the idea of a lineal progression toward the
given. Yet it easily becomes blind to the fact that by far not c2
everything belonging to its phenomenal region can be explicated ].5
completely in pure description. This is true precisely of those
analyses of phenomenology dealing with noetic-noematic clarifica-
Han of scientific modes of knowledge.
The difficulties emerging in this problematic area must be clari-
fied: the investigation will ha ve to begin with lived space, show that
the geometric spaces founded in it, and show how they
out of it in ATfn()'ligii ai first
ghinceNtliifi'"iiproich may- seem it
appears entangled when it is viewed by a thought that is required to
reflect u pon its own position. What must be kept in mind constantly
is that the subject, who faces the theme of space analysis in its entire
range of problems, can obviously only be a subject who is somehow
already in possession of all spaces. That mode and manner of
proceeding "from" the lived space "to" the geometric manifolds is
re-flexive in the literal sense, as determined by a retrogressive view,
does not present any special difficulties; in fact, it is the commonly
used method in phenomenological procedure. Yet this movement is
disrupted when we recall that the "way" of the subject through the
various spaces that we investigate is our own way, that the subject,
who is being viewed retrogressively in space, is in truth the observer
himself. While this revealing state of affairs may initially seem self
evident, its meaning must nevertheless be fundamentally rethought ,
in arder to surpass a merely lineal or progressive mode of phenom- V'
enological observation.
Thus it would be naive self-deception to "describe" the space of
sensory intuition while disregarding all determination of measure. lt
is indeed true, as will be explicated in a more precise sense, that
mathematical space has its foundation in the space of intuition. This
will be justified by our analysis insofar as it will deal with the space
of intuition prior to mathematical space. Yet as phenomenological, }
and mindful of a complete consideration of the "given," the analysis
must also mention that the space of intuition, in the only way it is
10 Introduction
) accessible, already contains aspects of mathematical space. After all,
the space of intuition is not merely a space of a sensibly intuiting
being; its investigation cannot be kept free from concepts whose
sense belongs to a "subsequent" context in which they can be
considered thematically and explicated more precisely. This is valid
mutatis mutandi for all "spaces" to be thematized. Thus the subjectl
does not traverse them as if they were a suite of
always be kept in mind that the subject, of whom we speak in the
third person for the sake of clarity and exposition, is in truth none
other than ourselves.
The investigation is placed in a difficult situation: on the one
harid, it must be to the factual necessity of proceeding
progressively from one spatial form to another; on the other hand,
from whatever vantage poirit has been reached, it is confronted by
factors that by their very nature can be made transparent only in
subsequent analyses. Yet such factors cannot be banned by any
decree from a particular moment of investigation. Our investigation
knows no other way to master these difficulties than by disrupting
the progressively advancing analytical work through self-reflection.
This would allow the incorporation of the results brought to the fore
into the next step of the process. Such a movement resembles a spiral
more than a straight line. The latter certainly remains in the
foreground of the work, yet it is not the sale constituent of our
method.
If by the term "phenomenological" one means all that is required
for the display and explication of phenomenal structures-i.e., the
"given," along with additional structures belonging to a specific
mode of access, and inclusive of the conceptuality required by such
a task-then our attempt can only conditionally be called phenom-
enological. The investigation intends to justify this characterization
in those parts dedicated to the descriptive and analytic work. Such
a point of departure, to be determined shortly with more precision,
will allow direct observation to speak for itself. lt will completely
avoid all constructions and derivations and will exclude all avail-
able scientific opinions "about" this area. This means no more and
no less than that the latter may .not function here with their
statements as necessary presuppositions; rather, they must attain
their meaning only within and "in the course" of the investigation.
Either they may be discussed for the sake of their own clarification,
or they may require phenomenological analyses to the extent that as
phenomena they belong within our framework. This would be the
case, for example, with geometry. This means at the same time that
Introduction 11
our investigation is exempt from any refutation by a theory that deals
with the problem of space constructively. Nonetheless, it remains
exposed to the possibility of missing the phenomena as well as to the
dangers of one-sided exposition of other opinions, and it obviously
remains open to critique concerning its own point of departure.
In what follows, the traces of Husserlian inflence will be manifest
only mediately. Yet the investigation is indebted to him even in
places where it is far removed from Husserl's phenomenological
position. Husserl himself had not thematically dealt with space
within the framework of his phenomenology. It was the merit of O.
Becker's work
5
to have applied phenomenology in an analysis of the
constitution of geometric objectivity. Our proposed investigation is
similar to Becker's in both theme and motive. However, our inves-
tigation differs from his in that it does not accept the Husserlian
brand of transcendental idealism as an incontrovertible assumption, .1(
as does Becker's.
To the extent that Husserl's own investigations-i.e., his analyses
of the spatial thing-have any significance here, they assume the
framework of his transcendental-phenomenological reductions and
serve to exhibit the constitution of a thing in pure transcendental
consciousness. Such analyses are based on the criterion of a process
of reduction whose sense and methodological correctness are based
on the display of objectivities in terms of the "how of their modes of
givenness" for pure consciousness.
For all that, the Husserlian transcendental reductions will not be
performed here. This does not imply a metaphysical presupposition,
one favoring realism. Rather, it primarily implies a
simple conviction that a dubious step can be avoided in Husserlian
transcendental phenomenology, a step that otherwise would be a
hindrance at the very outset to the mastery of the problem of space.
While the radical accomplishment of his reductions also leads to the
reduction of the factual-empirical ego to pure consciousness, we
shall counter this by pointing out that the subject, strictly under-
stood, resists in principie any reduction in the Husserlian sense. I.he .
subject must be reestablished in his full and concrete su]Jject_iv:ity ...
i.e., in Qisf(1SJ2!y .. iifll_,
for the phenomenology of It is justifiable and indeed neces-
. '''' --:..>-0'"''""- -
5. O. Becker (1). Numbers in parentheses after an author's name refer to
the works in the bibliography listed by such a number under each author's
name. Only the frequently cited works of Husserl will be given with the
standard abridged title for the sake of convenience.
12 Introduction
sary to exclude all that is "merely" factual and to disregard all that
is conditioned situationally by the individuallived body; neverthe-
less, it must be maintained that to neglect facticity and corporeity
entirely would mean a premature move toward metaphysics and
away from the methodological correctness of descriptive efforts.
Thus the subject after the reduction would not be grasped phenom-
enologically as a subject, since in Husserl it functions in the sphere
of pure transcendental consciousness as constituted in accordance
with the requirements of a spatial thing. Toward the end of his
works, Husserl noted these and related difficulties and recognized
the subject as something that cannot be included in the process of
reduction in the form in which the process was initially inaugurated.
Our arguments against Husserl are not in tended to be a destructive
critique of his phenomenology; due to an entirely different point of
departure, such a critique would be fallacious. A transcendent
position can serve only to confront two theoretical stances with the
aim of clarifying their own thoughts. Nevertheless, a critique of
specific Husserlian points is justifiable when we can take exception
to mistakes within his phenomenology, and specifically when such
mistakes could have been avoided in terms of his own point of
departure. Husserl himself would not have denied us this right.
PART ONE
LIVED SPACE
SECTION ONE
Contributions to the Phenomenology
of Lived S pace
Point of Departure and Statement of the Problem
The point of departure of this investigation is prior to any decision
J concerning the independence
1
of the external world, reality, or
being-in-itself from the subject. It begins with the observation of a
form of consciousness that as "natural" or "everyday" conscious-
ness it is nevertheless far removed from being realistic. Its position
toward the world is nothing other than one of being straightfor-
wardly given over to the world, of direct respect for the world in its
demands and assurances. Its statements contain neither realistic nor
any other kind of "istic" meaning; only reflective thought is exposed
to the danger of interpreting them in such ways. To grasp entities as
beings-in-themselves oras merely subject-related is to presuppose a
reflective attitude. This attitude requires that the subject extricate
himself from the stream of events and experiences and differentiate
himself from the world. However, the ontological sense of prereflec-
tive expressions is metaphysically indifferent. It consists in nothing
other than an immediate being "by" the things and being "in" a
world.
A spatial relation is therewith already expressed, although the
illumination of its nature is left for later pages. Here it cannot yet be
assumed in advance insofar as the subject, still to be considered,
does not yet grasp himself with respect to the meaning of this
spatiality. For him the living being is justas spatial as are things, and
like them, he too exists "in" space. If asked about the nature of space,
he takes it to be something empty that appears filled with worldly
things, events, and states of affairs, and thus to be simply a "world
space." Yet su eh a conception arises from a theoretical, even if
primitive, attitude; strictly speaking, it is an answer to an explicit
14
Contributions to the Phenomenology of Lived Space 15
question, but not a primordial relationship to space. To test this
answer in terms of its meaning and justification is reserved for
subsequent discussion. Of crucial importance here is only that such
a "position," based on a response to an explicit question, is clearly
removed from all immediate otientation to the world. In this latter c9W
case space is not thematized. It is pre-reflectively there in the
process of corporeal and intellectual activities without becoming an
object for consciousness.l
Our investigation ought to begin with the analysis of this pre-
reflective world-posture. The subject must not be interrogated pri-
marily in terms of his judgment "about" space, but in terms of his
comportment "in" it. Of course, the subject does relate to space by
making judgments "about" it. This constitutes his very being as a
subject: to know oneself in "opposition" to the world, to have space
objectively and to be able to express judgments about it.
This ambivalent awareness of space obviously plunges the inves-
tigation into inextricable difficulties at the very outset. How can it
adequately grasp the relationship of a being "in" space when this
being is essentially constituted by being "over against," outside of,
space?
The formulation of this question reveals at the same time another
problem. Even the notion of "over against" suggests a spatial
relationship, which, as one is apt to say, is nevertheless meant
non-spatially. A brief reflection on language readily shows that
thinking, regardless of its' efforts to escape the power of spatial
metaphors, can express itself only by succumbing to their
misleading force. This is not only the case when thought employs
them deliberately for the sake of vividness, but even when, with full
insight into the inadequacy of such images, it attempts to exclude
every spatial meaning transmitted by a word. It would require
far-reaching philosophical explications of language to develop the
problem of space from this viewpoint. Here we must forego such
considerations for obvious reasons. We are excluding this type of
semantic reflection and shall take extreme care to provide the
greatest possible clarity and terminological univocity.
1. This is why we speak here of lived and not experienced space,
specifically when the concept of experience is used in phenomenology in an
almost unlimited terminological sense without being appropriate for our
subject matter. The concept of lived space was already used by K. v.
Drckheim in 1932; however, our investigation follows different method-
ological principies and thus results in different formulations.
b lA-\ O.t
prv \vlr-\....


JY
16 Lived Space
The return to the question raised abo ve concerning the proper point
of departure requires a closer look at the constitutive relationship
between the subject and space. It is based on the subject'smode of
beingas a corporeal subject. 'I'he lived body must be understood in Le-
a strictly phenomenal sense. This means that it can be taken neither
asan organism nor as a physical body infused with a soul. As special K&
sciences, as methodologically established arrangements for research- .l
ing corporeity, physiology and psychology imperil the view of what \..et
is given directly and originally: primordially, corporeity is neither a
system of organic processes nor a body "inhabited" by a soul, nor
even a "unity"-regardless of how conceived-of body and soul. It is
also not primarily one's own lived body with specific sensations or
inner content ("states"); immediately and originally, the lived body
is presentas another living being and, more specifically, as the com-
portment in which this being's relationship to the surrounding world
is announced in both a sensorial anda senseful way. The lived body
always appears as lived body only in such comportment in a situation
and is understandable immediately from the situatwn; this
mode of understanding is prior to all scientific explanations. No
analogical inferences, empathy, or other auxiliary constructions of a
sensualistically oriented theory used to grasp the other are of value
herei Rather, corporeity is _ _given in a way that is indifferent to any
regional distinctions between the QYChic and the 12hysical, the inner
andthe outer. also quite accessible to the
either aspect- even if the conceptual structure of
this "also" need not concern us here- we must maintain a strict
distinction between the corporeal phenomenon given in immediate
presence and corporeity as determined by specific methodologies.
2
2. The anthropology of the twenties has insistently pointed to this
"psycho-physical indifference" of corporeity. See M. Scheler (2); see also
F.J.J. Buytendijk, H. Plessner (1), (2), andE. Rothacker. This anthropological
research has served well to rediscover an aspect of corporeity overlooked by
the special sciences ever since the Cartesian dualism of body and soul. If
anthropology insists that its position is prior to all specific researches,
nevertheless the ensuing conception of the neutrality of its point of
departure cannot be construed as indifference of philosophy toward the
individual sciences. In its own progress, philosophy must interpret scien-
tific results with its own possibilities of reflective thinking; it must la y open
their sense by introducing all the methodological precautions of the indi-
vidual sciences, which in their turn-as specific "aspects"-require
structural elucidation.
In the latter aim, anthropology is distinct from existential philosophy.
,....,.., .,_...,.., ..
a f\ eeJ .... \ur w
Contributions to the Phenomenology of Lived Space 17
Corporeity is understandable only frorn its cornportment toward the
world, and it lends itself to phenomenological observation only
when approached in terms of differing styles of its relationship to the
surrounding world. We can indicate three such styles: asan attuned
corporeity, it is a carri_er of expressive content; as a practica!
corporeity, it is a point of departure for goal-oriented activity; andas
a unity of senses, it is the center of perception. L
01
b e,kJ
It will become apparent that each mode of being of the corporea\
subject corresponds to a specific spatial structure, or, to be more
exact, that "the" one space is obtained, structured, and filled
differently depending on the cornportment of corporeity. It is quite
clear that we are not dealing with three separable forms or three
temporally and genetically distinguishable spatial steps or stages, as
if the homogeneous space of an objective space-consciousness were
capable of being deduced or developed from them. Rather, this
singular space presents a contingent presence in objective con-
sciousness that is completely unavoidable, given our point of
departure. The point of departure must be accounted for, although
its elucidation must remain a matter for subsequent considerations.
The following investigation of the three spaces is concerned only
with this: to elucidate and present the variously structured ways the
subject, as corporeal space in accordance with the ?
various corporeal
It is important to note that in this context the differentiations are (
ontological. The corporeal modes of comport-
rnent are not psychological. Although there is a psychology con-
cerned with the genesis, structure, and performance of expression, of
pragmatic processes, and of perceptual activity, this does not mean
that ontological questions concerning the nature of expression,
action, and perception and the understanding of their whence and
whither are superfluous. In fact, it is assumed that such questions are
already answered insofar as they provide clues for the proper
interpretation of the results of individual sciences. Modes of com-
portment are here not conceived as objects of investigation of an
While the philosophical by
in French existentialism the lived body assumes a
preeminent place as a theme of exhaustive structural analyses (see the works '52:
of J.P. Sartre and M. Merleau-Ponty; see also the survey by A. Podlech).
Their emphatic rejection of all efforts by the special sciences provides an
occasion for critical reflections to be offered later in an appropriate place.
)(, 1..-vC\<-"',;o;"
vJ ~ ( \ v J
v{- ~ \ ( _ W<\t
18 Lived Space
extremely broadened theory of behavior, but rather as a kind of
senseful relationship of corporeity to the world.
The differentiation suggested above is not episternological, just as
the relativity of lived space with respect to the corporeal subject is
not epistemological relativity, but a relativity of being. The three
emphases suggested have to do with the modes of being of givenness,
not with modes of givenness of entities. The latter might assurne that
within a particular position toward the world, entities could be
given otherwise than their type of being would require. This is
nevertheless excluded by the strictly correlative relationship be-
tween world and cornportment toward the world . .Thus in the
attitude of expressive understanding there are no things given "for
the sake of .... " Where expressive understanding suddenly "breaks
forth," then, taken frorn the side of the things, the thing is "given"
ontically otherwise than befare; yet the transformation of the rnode
of givenness is ontologically grounded in an altered orientation of
the subject. While such an orientation is comprehensible in terms of
an entity of a particular kind, the entity is only understandable frorn
the subject's comportment.
The relationship between space and subject cannot be sirnply
accepted as a reciproca! relationship without further reflection. The
old, enrooted-c-n.c;;ption of receptaculum rerum, assuming space to
be identical and indifferent with respect to the changes of its
contents, is contrary to such a relationship. A more detailed expo-
sition of this will be offered subsequently. With the proposed
division into three styles of corporeity, the investigation is forced
into differentiations that are justifiable only rnethodologically.
While our investigation separates the attuned space frorn the space
of action and the latter frorn the space of intuition within the lived
space, according to the three modes of corporeal cornportment, in
arder to show their characteristics and properties separately, we
remain cognizant that we have assumed separations and divisions
that are valid only if the analysis, with its increasing precision of
distinctions, is capable of adding to the clarity, visibility, and
cornprehensibility of the unity and intimate relatedness of what is
being analyzed.
Chapter One
The Attuned Space
Jl-v wt-k u.,M_
1. The Concept of Attuned Space

The attuned space is hardly accessible to conceptual thought.
Certain misunderstandings that could make such thought more
difficult must be avoided at the outset.
W e must first defend our thesis against the conception that space
is gr ven ly
a mere medium of
1
(Q)
measurement. In its most primordial ontological form, space is on -:>--
side of numerical and quantitative determination; its ,
primary characteristic lies in its being a quality and an expressive
fullness. Grasped in its unique immediacy, it is all enveloping and
Cinpromises the "atmospheric" dimension of an attuned being. It is
a space of labor, of leisure, of festivities, of devotion-a space that is
loved, hated, feared, and avoided. As a medium where human life is
realized, it has its own proper visage. It means calamity or seclusion,
a foreign place or a home, a place of transient residence or an
enduring stay. It is different in accordance with the differences in the
being who it. W '-" /)
The this space is not peri:eptwn, a.gd awareness 'tx-\-f<?f{.ettt-1.
of space is not cognition; it is rather a way moved and
affected. S pace indeed exercises an "effectivity," yet its relation to
1
W,
11
experience is not casual; rather it "addresses," "imparts." Space is
not primarily an object for a subject who performs acts of spatial
understanding. Rather, as space, an
of coexistence Such coexTsfence escapes all the
1
conceptua.Tdeterminations of a thought founded on the opposition of v'
object and subject as a "relationship" or "connection." All these in
their turn are founded on the primordial and intransgressible bond (}
between Thus lived experience here Sr
does not mean an oriented engagement in the sense of act phenom-
1
<oroTu{ ?
20 Lived Space
l
&- 't d' t' . h . lf f
eno ogy; as an ai uned llved' expenence, 1 1s mgms es ltse rom
oriented engagement by a lack of intentionality. Neither does it mean
a corporeal state within the subject; thus, strictly speaking, the
concept of sensation must also be avoided. Here lived experience
means a unique communication of the living- experiencing ego with
another, with an expressively animated space.
3
A further objection must be countered; namely that the attuned
space emerges only during a particular temporal phase, during the
"exuberant moment." When our subsequent discussion occasionally
touches upon such a moment, it will be for the sake of exemplifica-
tion, as an appropriate and powerful manifestation of salient fea-
tures. Yet such features are not restricted to these examples. Expres-
sive understanding is a unique mode of orientation toward the
world; expressive understanding has its own sense-context, and it
must be taken in this uniqueness and interrogated with respect to its
space.
This does not contradict the notion that "spaces" may appear to be
different. The transition from one to the other, from the space within
a church to the "animated" street, does not imply a dissolution of the
atmospheric space as such, but rather merely a change of the
expressive content. The difference between such "spaces" is itself a
- __ ..___ _____ .--------=.,. __________
positive determination of tb. attl1:Qe(LSI@_G!Lfl.S such.
---------- ----
Expressive Understanding is also not contradicted by the fact that
it is disregarded when the attitude of the subject is determined by
practica! or theoretical aims. Rather, having a purpose makes it
necessary for one to perform a specific shift, or-to speak metaphor-
ically-to step out of the attuned space into a space with a com-
pletely different structure, consisting of goal-oriented activity, of
sensory intuition, or of pure thought.
Attuned space is encountered in a pre-reflective orientation to-
ward the world. Such orientation to space is an immediate affinity
with the world. For an attuned being to be in another "space" and to
live in another "world" are expressions identical in meaning. They
testify, on the one hand, to the fullness of sense and meaning
characteristic of the space discussed here and, on the other hand, to
the specific difficulties involved in the conceptual determination of
attuned space. While at first the attuned space is bound to limited
spatial surroundings, it appears to be capable of multifarious exten-
3. Concerning the concept of attunement and its relationship to feeling,
see O.F. Bollnow, S. Strasser, and P. Schroder. The concept of sensation will
be discussed subsequently in greater detail.
The Attuned Space 21
sions. We speak of the space of our future and our past, of our wishes
and our hopes. Home and away are not only significant wholes, with
appropriate feeling and mood accents; they are at the same time
understood and differentiated spatially. Within them the human also
]
"moves," and things and events have their "place"; they are "near"
or "remate," getting in our way or lending us a free "way."
Seen from the vantage point of objective space, all these are
obviously mere metaphors, spatial images for relationships that "in
reality" are not spatial in kind. Yet this objection does not touch the
modes of ex eriencing such spaces; after all, our lives within them V
are not merely imaginistic. The objection cannot explain why space
plays such a preeminent role in lived experience and why specific
unities of meaning offer themselves spatially to lived experience. A
more exact investigation in this direction would lead us too far afield
from our present purposes. We shalllimit ourselves to attuned space
in a narrower sense.
4
4. It will be shown that this space includes temporal determinations. This
relationship between space and time is stressed by E. Straus (2)-at least
through an application of an ambiguous and misleading concept of sensa-
tion. Yet the relationship between space and time should not lead us to
encompass everything within attuned space, within whatever is designated
by "space" in the metaphorical use of the word. Phenomenologically, a clear
distinction must be made between the attuned space in a genuine sense, i.e.,
between a space given as "real" under specific and always freely actualiz-
able changes of attitude of the subject who can find himself in it as real
corporeality, andan attuned "space" in a transferred sense. The "transfer,"
in its turn, is a clearly demonstrable datum of lived experience; even when
we "live" in such spaces and have "resettled" ourselves in them, then, given
a corresponding attitude, the attuned space can become in principie a
specific datum of lived experience. This in turn is possible on the basis of an
already understood spatiality in the genuine sense as defined above.
The complex of phenomena constitutes a unique field of interesting and
specific analyses wherein it is possible to distinguish various types of
transfer. Here is not the place to investigate them further. Only different
metaphorical meanings of space should be mentioned: they are found in ?--
~ i o m s , in----peiry,nthe illusory spaces of drama and their
numerous realistic-spatial surroundings on the stage space of the theatre, in
the imagined and imaginable spaces of fictionalliterature, etc. It is remark-
able that the contemporary theory of literature pays specific attention to the
theme of space (see Bachelard, Blanchot). The motivation for a typological
reevaluation of the literary aspect of space stemmed from W. Kaiser. This
dissolved the traditional principies of articulation in the theory of fiction,
principies that previously pertained purely to the subject matter. See Kaiser
22 Lived Space
2. Characteristics of Attuned Space: Fullness and Emptiness
The primary access to attuned space is offered by the character of
Jhings_l!0.J;. This does not constitute a prejudgment concerning the
relationship between space and spatial things. No matter how this
relationship will be determined, it is in any case clear that space as}.
such is accessible only from its fullness. It is not readily obvious at{
the outset whether it is possible to investigate an empty space. What

from the spatiality of things. The characterization of space is
only through a delimitation of
various kinds of thingly attributes. But the notion of "thing" in
attuned space must be taken with a grain ofsaTf u pon
it, leaving our thought with its customary although insufficient
categorical distinction between things and properties. Yet strictly
speaking, it is precisely in the lived experience of expressive things
that such a distinction does not occur .
and object, _ which _ thj_s_ disti:qgtion presupposes, is a __ recent
thng as
lav prperties be perceived; .with - its
"character."
5
It is no accident that our attempts to describe the lived
--
experience of expression lead us to specifications suggesting, and
drawn from, the psychic domain. This implies neither a misuse of
language nor an inappropriate anthropomorphization, even when
we do not take things in their objective color- or form-
characterizations, but rather in their own unique and momentary
"toning." Exuberant or sad, hard or tender, soft or severe-these are
both sensory formations and sense-formations, symbols in Goethe's
sense, which lend us "specific dispositional moods" anl which at
the same time possess "sensible, moral, and aesthetic purposes";
one can "even employ them as a language when one wishes to
express primordial relationships."
6
Goethe's hesitation, his suspi-
cion of being exposed to the ecstatic, seems to be baseless. The lived
(1), pp. 360-65, and (2), pp. 24ff. See E. Stroker for an entirely different
typology of imaginative space in painting.
5. The distinction between property and character was first made by L.
Klages. Similarly, J. Konig, following the fourth of E. Husserl's Logical
Investigations, differentiates between determining and modifying predi-
cates (pp. 1-16). For a characterization of the latter and its specific kind of
relationship to the subject, see especially pp. 32-41.
6. J.W. Goethe, Sixth Section.
The Attuned Space 23
experience of attuned space expresses itself readily in such primor- S
dial relationships. Prior to objective recognition, its forms are 7
already understood "physiognomically."
In attuned space, therefore;teasfinction between primary and
secondary qualities does not exist. The form of things communicates
as expressively as does their color. Both are equivalent in their
relevance for mood; in their physiognomic content they are capable
of mutual support, enhancement, or dissolution, and of eliciting
dissonant experiences. Even the size of things is here far from being
a mere quantity. The experience of the powerful or the sublime, as
well as the graceful and elegant, belongs essentially to size. It reaches
complete fulfillment where its harmony is sustained and maintained
with the rest of the characters. Regardless of how it may be
determined, regardless of what aesthetic criteria may be sought for it,
it is decisive that even size remains incorporated as a moment of the
expressive totality, and thus it cannot be extricated and given an
independent character without the loss of its valence of mood. The
of .. _is .
espec1ally convmcmg m a trend of p1ctonal art that mcorporates m
0
'1t
its formation the effectivity of precisely these two momnts.
"Expressionism" owes its expressive power, not to mention its
sovereignity, to the presence of form and size factors not bound to
any objective attributes of a thing.
Form and size are constitutive factors by which we apprehend )-, . +,
generally n_ot o_nly the perspe?tivity of space: als_o all the rest of
the determmatwns related to It, such as centnclty, onentedness, and
finitude. What can be said about them with respect to attuned space? n
By raising this question, the subject betrays at the same time the fact
that he already belongs to another space; no other state of affairs
underlies the claim that the spatial perspective is conceivable in
terms of the size and form of things thaii that the size of things is
observed comparatively as greater or smaller in relation to others and
that their forms are consciously perceived in certain truncations and
intersections with others. But this means that they must be regarded
as purely quantitative determinations. As such they make their first
appearance in the space of intuition. is ')
tion. If we infer from this that "there is" no perspective in attuned
space, then we require additional explications in arder to preclude
closely related misunderstandings.
Indeed, in attuned space things are encountered in a specific order
of being behind one another. Yet while size and form vary in
accordance with perspectivallaws, the constitution of attuned space
24 Lived Space
does not depend on them. That the perspectiva! arder can become
physiognomically significant and participate to a great degree in
determining the spatial atmosphere is not contested here; this is in
accord with the functions of form and size as expressive characters.
Thus it is sought in pictorial presentations of this space for the sake
of its mood-bearing moment. Yet it is essential to incorporate this
phenomenally incontestable state of affairs appropriately. Attuned
space "is" perspectiva!, but not because perspective is a structural
property of this space; the claim is rather that the lived experience of
expression evoked in attuned space by the perspectiva! arder of
things can only be given because the experiencing subject in this
space does not live without the objective space. While living in
attuned space in a most primordial world-posture, prior to all purely
objective orientation, the subject already has the objective space
"behind his back."
Attuned space therefore bears determinations that allow it to
appear as profiled against the background of the pure space of
intuition; thus it is never free from the determinations of the latter.
Even the perspectivity of attuned space is a characteristic present by
virtue of the space of intuition. Yet perspectivity does not belong to
attuned space as such. The resulting consequences for the orienta-
tion1and centering of attuned space will be pursued in subsequent
paragraphs. Meanwhile it is essential to touch upon a phenomenon
that, along with the color, form, and size of things, possesses great
power with respect to atmospheric space.
Sound plays a uniquely significant role in attuned space. In
space-determining power it surpasses even color and form. At first
this appears dubious. After all, the arder of colors and form is an
arder of one next to the other, i.e., the genuinely space-constituting
arder, while sounds are present one after the other in time. More-
over, color and form can be more precisely located in space than can
sound, which only indicates a direction.
To counter the first doubt, it is necessary to become free from the
presupposition that every space must be defined as an arder of one
next to the other. It is not yet certain that the space investigated here
can be circumscribed by the classical Leibnizian definitin. If the
claim about the importance of sound could be confirmed, then there
would be an indication of a structure entirely different in kind.
Although it is a matter of gradation, localization is attained with a
differing precision for the colored object and for the source of sound.
In contrast, between color and sound there is a complete qualitative
difference that touches upon the spatial characters of both.
The Attuned Space 25
The color is attached phenomenaliy to an object, not only as a lc\or / ~ u
property, with all its nuances and differentiations of intensity, but
also asan expressive presence. Color can never appear except on a
colored object. It is otherwise with sound. Sound detaches itself
from its soun;e. It is not a property but an event; it is not attached to
something but draws nearer and recedes into the distance. It can
indeed be said that it is a sound of something, that points to a source.
But this character of originating with something is, strictly speaking, Souttdjflb,'i
not appropriate to sound but to noise. Noise is always noise of
something and is always perceived as such. Sound, in contrast, has
an existence that is detached from its source; it becomes a sound
precisely because it is capable of such detachment. This state of
affairs reaches its completion in music. Here it is the sound itself that
fades away and strikes up, not the cello or the first violin. Of course
sound can be heard as the sound of an instrument; but in this case we
are dealing with an entirely different mode of lived experience. Not
only does the expressive content attain a purer and more complete
presentation and effectivity in the lived experience of the pure tone
formation, the sound experienced in its free and appropriate exis-
tence has a different spatial character. It will be experiencel more
intimately than a sound of an instrument; it pervades space and
determines its atmosphere more completely than it would if it
were perceived in relation to its source.
The same process of detachment can be traced with extreme
clarity in spoken language. It too is primarily a sound formation
brought about by specific organs of the body. Yet the experienced
sound is not merely the sound of a speaker. By virtue of the sense
content of the word, the detachment occurs here so completely that
one requires a specific shift of attention in arder to grasp it as a word
of the immediate speaker. This shift will be spontaneous when
someone speaks a language of which we "do not understand a
word." Here the word will be come noticed as a word spoken by
someone, who emerges into the foreground with all his personal
accentuations. With the emergence of the shift of signification from
the spoken to the speaker, the experience, in its conceptual sense,
approaches noise. As a sensory formation, the word is subsumed
under laws, making possible a theory of language, just as sound
obeys the laws of harmony. The clatter of a motorcycle, however,
does not follow any musical syntax. Noises are not objectifiable; ~
their "sense" exhausts itself in being noise "of something." This is
not the p ace to ela orate furhher relationships and differences
between tone and linguistic s o n ~ d. We have memly suggested how
Cli ~ J
+/5
26 Lived Space
the sense-fulfillment of a sound formation goes hand in hand with an
emancipation from its source.
What sort of determining forces noise and sound have for space
becomes clear when the same space is experienced once as sound-
less and then as filled with sound. Colors and forms maintain the
same order and distance, remaining remate forms that change only
when dusk and night blend and dissolve their contours. The sound
i
coming toward us not only fills also contracts it. A distant
church moves closer wiTha resounding of its bells. lts inner
space becomes noticeably smaller as soon as it is pervaded by the
sounds of the organ; in a siient film, the screen attains a peculiar
distance.
A relationship of this kind between space and sound presents a
new problem. The sound formation taken in itself is a temporal
formation. What remarkable relationship exists here between space
and time? We shall return to this question in another context
(pp. 36 ff.).
The qualitative content of attuned space is not exhausted by being
an expressive content of color, form, and sound. lt is possible to
suggest a domain of lived experience in which the experience of
expression has remained in its purity. This purity must have
persisted not only across the objectified and reified world, but has
remained inaccessible in principie to any other apprehension except
the physiognomic. Literally speaking, animate creatures also belong
among the "things" of the attuned space; they constitute its atmo-
sphere more clearly than do inanimate things and their spatial
valence is easier to detect. From the standpoint of objective space-
consciousness it must be complete! y inconceivable that ill. J)}.I'J' . .
presence "space would become immense" for us, or "too

'(i
that they could allow us "leeway" or could squeeze us out of our
"place," compelling us to leave our location and to move away from
them spatially. This sphere of the expressive experience of fellow
creatures shows most clearly the remoteness of attuned space from
all measurable determinations. lts distances are other than measur-
able. the concept of. the
region of the measurable, then g_cag_be said that there are no
attuned space. ---
----Tlirough-Itsfil1ness, attuned space has become visible, at least
allusively. A question could be voiced concerning with what justi-
fication we infer space from fullness-what might space otherwise
be? The sense of such a question implies that space is something
other than its fullness. lt assumes an empty schema capable of being
The Attuned Space 27
seized in itself. Perhaps this question js influenced by conceptual
analogies of the space of intuition. Yet despite the manner in which
the relationship between space and thing may be structured in
intuition, the essence of attuned space excludes the differentiation
between empty and full space. To wish to extricate an empty form of
attuned space from its formations would be senseless. Indeed, it too
contains a lived experience of emptiness, the encounter of "noth-
ing," the dissolution of its formations, which "have nothing more to
say"; yet this emptiness, experienced as a mood, is not to be confused
with the empty form of space sought by abstract thought. Attuned
space is not merely given and lived in its fullness, but rather it is this
fullness itself. Its loss is the loss of attuned space. This is the reason
why we speak of fullness as we would of attuned space itself.
3. Place and Po sitian in Attuned S pace
Attuned space has its general characteristic in being a form of
expression. As such, it is nota manifold of positions, nora system of
dimensions. Place and position in it are not determined by an
assertion of quantitative relationships of a there to a yonder: Even
m y own place in it is never merely a point determined relative to the
place of things. I am in my workroom-this does not mean being 1
somewhere "next to" the chimney. 1 allow the inside of a
Romanesque church to affect me - yet my "place" in it is not at a
comer of the quadrature. Certainly, these positions belong to the
space in which 1 reside; its formation, its architectonic articulation,
are part of its fullness and shape its atmosphere. Yet they do not
determine my place in it; for my lived attunement, no tangible
relationship holds between the lived experience of space and the
position in which 1 find myself as a corporeal being. M y phenomenal
place in attuned space is not ascertainable. As an attlineabeing,I
have no determinable locatioln

Thus in attuned space my lived body cannot be a point of
reference for a relative determination of the position and the place of
things. Attuned space has no center of reference from which it would
be possible to arder and separate the experienced things and
determine them as there in relationship to a fixed here. lndeed, in
m y vision, hearing, and touch, things are constantly present and thus
are related to me in my corporeal organization; yet this relatedness
not only does not play a part in the actual experience of things but
even in reflection on experience it does not appear as a facet of
attunement. It is a relatedness of things as pure perceptual objects to
28 Lived Space
me as a corporeal-physical body, and this relatedness as such
belongs to another space. Relationships from the oriented space of
intuition reemerge here quite readily. Yet here this space is not yet
thematic; in attuned space it remains veiled in the sense described
above. Its perspectivity is present in attuned space without being
part of attuned space. That attuned space does not possess a center
must be taken in the same sense. The centricity of space requires its
relationship to a corporeity that is necessarily of a type that is
localizable, i.e., a corporeal-physical body. Indeed, as an attuned
corporeity I am not without location, yet it is not through location
that 1 am an attuned corporeity. Corporeity as a bearer of expression
has no spatial position.
Lack of orientedness is closely related to the absence of a center in
attuned space. More precise results will appear in the analysis of
expressive movement. In the attuned space "there are" no preemi-
nent directions to be taken in arder to attain something, as would be
the case in the space of action. When one follows directions of roads
on a map, one does so in arder to orient oneself-yet one does not
orient oneself in the mood space of the landscape.
Attuned space yields itself fully and completely first of all in a
purposeless lingeri;g
there Tii- spac ,-ai:bifiarlly and
externally interchangeable. The thing, as a bearer of expression, has
"its" place. The place belongs to its expressive power and accentu-
ates or diffuses the physiognomy not only of the thing but of the
space as a whole. An exchange of place of two things not only
abolishes their genuine expressive fullness but also noticeably
)
disturbs the atmosphere of the e.ntire spac.e. P. ieces do
(]!) not "belong" in the marketplace, nor church windows in an office.
7
It is precisely
clashes of style, that the place of the thing appears most strikingly;
on the one they reveal that __
-;
appropriate relationship is capable of transforming attuned space as
a whole and of disrupting its atmospheric unity. This shows clearly
how little it can be characterized as a mere proximity of places and
locations. Its structure deviates markedly from a mere manifold of
positions. Rather, it is given as a closed unity of expression that can
7. K.v. Drkheim, p. 406, already indicates these lived experiences of
"belonging there."
The Attuned Space 29
be affirmed or negated, accepted or overlooked as an inseparable
whole.
4. Nearness and Remoteness
Sin ce in a
iJLYiilid.iOL.its ... d i sta nces. In
attuned space there are no measurable distances. This is understand-
able from the qualification of things by place and from the holistic
structure of attuned space. In.metric space, distances between things
can be increased or decreased without incurring the same changes in
the things themselves. In contrast, in attuned space such changes do
involve the things themselves: their "in-between" is not a mere
relationship of arder to be considered apart from, and in comparison
with, other things; rather, it is a qualitative, expression-bearing
characteristic of the thing itself.
As such, it is related to the "place" of the experiencing subject,
i.e., to his being attuned here and now. The nearness and remoteness
of things open up to the subject. The difference between them is not
a mere matter of degree; nearness and remoteness differ qualita- <J
tively. A distance is composed of smaller segments. Yet nearness
does not consist of other nearnesses, and remoteness is not com-
posed from various remotenesses. Distance can be separated into
partial distances, but remoteness cannot be divided into remote-
nesses and nearness, into nearnesses. A remoteness does not contain
another remoteness, just as nearness does not contain another
nearness and an addition of nearnesses does not result in remote-
ness.8 Distances subsume things as perceptual objects, but not
nearness and remoteness. The latter are expressive phenomena,
although expressive in another sense than color, form, and sound.
The nearness of a thing in attuned space does not have its meaning
only because its content appears clearly, but because it is near tome.
Nearness and remoteness are not attributive determinations of the S
thing grasped, but of my mode of grasping. J
Nearness is pure being-present, lingering here and now or being
threatened or beset by things, which, by their irritation, intrusion,
and oppression, leave no "space" for one's own relations and
movements. This space can be established in two ways: in yielding,
fleeing, and removing oneself; or in surpassing, overpowering, and
8. In another context the irreducible characteristics peculiar to nearness
and remoteness are mentioned by E. Straus (2), pp. 288ff.
30 Lived Space
overcoming. The constitution of the remoteness of attuned space
takes place in this dual movement; thus remoteness is that which is
"no longer," it is "there" where I no longer am, where I have ceased
to be present- a limit of my attuned space. It lies "behind" me, it
has departed from my view and vanished from my actually present
lived experience. Yet it is also something else: what is "not yet,"
what is still veiled from my view, and sois also a limit of space that
lies "befare" me. It becomes an aim of my search, an orientation of
my departure. Nearing is not the dissolution of a remoteness.
Distances can be traversed, yet remoteness can never be reached.
With each nearness appears a new remoteness. It is only nearness as
an attained here-now in a coming from a no-longer and lingering
befare a not-yet, just as remoteness is only the surpassing of
nearness. Both are reciprocally related.
It is striking how spatial and temporal relationships pervade one
another here. Nearness and remoteness are spatio-temporal phe-
Q nomena and cannot be conceived without a temporal moment.
While yielding a horizon for attuned space, they also contain
another aspect of conceiving this space as "space-time." It is only
where nearness and remoteness are abolished-i.e., in metric
space-that space and time are sundered. Distance is a purely
spatial quantity.
------
5. Movement and Orientation in Attuned Space
Nearness and distance are relative toa motile living being capable
of approaching or distancing itself. It may seem that the problem of
motion is secondary iir significance. Since all motions . occur in
space, then obviously space must be assumed as the "play-space" of
motion. Hence the analysis of motion could be avoided for the
understanding of space. Space appears as a precondition of any
possibility of motion.
At the outset, the following investigation will be based on the
phenomena pure and simple, in a specific sense-i.e., we shall
observe the phenomenon of the self-movement of the subject.
Self-movement always served as a sign of the living: if something can
move "itself," can follow its own impulses and not mechanistic
causes, then it belongs to the living region. To the extent that life
articulates and unfolds itself in its richness of gradations, it also
grows in the richness of its forms of movement. The richness oflife's
growth and the increasing development of movement variations are
The Attuned Space 31
mutually implicatory movements of the same unitary development.
9
The bodily dynamics distinguish themselves from mechanical
motion by more than the vitality of inherent drives. Were this the
sale distinguishing characteristic, then the perception of movement
would be reduced to an organism traversing a basically meaningless
i
series of linear points. But in fact each apprehension of a bodily
movement is constantly transcended toward a specific intention: the
movement is a searching, defensive, furious, tired, happy movement
and hence is already grasped as a dynamic mode of relationship to
the world. It is understood from within and in relation to its
situation; both limit one another. Thus the question as to whether a
pre-given situation motivates the movement or whether the move-
ment constitutes the situation cannot seriously be asked.
This means that expressive movement unfolds itself and is com-
prehensible in its total fullness only in and from the space wherein
the attuned being lives. That one moves differently in a
than in a factory:, can
be understood solely by understanding the coordination of reflexes
ano muscularccmtfficfions. Tlis is not to deny the fact that such
coordii1at1oon1leinovient of parts may constitute a necessary
condition. Nonetheless, expressive movement must be grasped from
the experience of the space "in" which it takes place. But what is the
sense of this being-in?
In its singularity, attuned space is always a specific space of
motility. It allows and requires specific forms of movement and
excludes others. The movement of my lived body is a comportment
toward the expressive content of these spaces, and it either fulfills or ()
rejects their demands. At the same time, it is a space of movement in
another sense. If my attunement is created by the atmosphere of
space, then my movement appears to be formed by it and attuned
toward it; yet this is only one side of a reciproca! relationship. There
is also the other side. Space is not just a space
becomes a space, in its specific attunement through my movement.
Through m y movement 1 canegafe-If, atfeSlTo1T;'O'fCl'se mys8Ifoff
9. The contemporary theory of science of biology and physiology and its
"introduction to the subject" (J.V. Uexkll, A. Portmann, F.J.J. Buytendijk,
V.v. Weizsiicker) has freed itself from the path taken by a purely physical-
istic approach to physiology; it explicitly sees itself as an effort of "under-
standing," without devaluing the results of earlier researches. The work of
Buytendijk, in particular, is outstanding in its strictly delineated conscious-
ness of method.
32 Lived Space
\
from its content. Thus space no longer remains what it was, but is
immediately transformed. In a reverse manner, I am not only a
receptacle for its contents but a co-carrier, and first of all a shaper of
its atmosphere through my movements.
Attuned space is not just an expressive form "within" which my
movement occurs. lt arises as the space of my own movement. My
expressive movement is exactly what its spatial content is as mine;
in turn, the spatial content itself has a moment of my activity. If we
keep in mind this strictly reciproca! relationship between space and
expressive movement, then we have a principie for attaining a more
detailed exposition of the structure of this space. lt was already
noted that this space is not a pure manifold of positions, but a
formative whole. The phenomenon of movement allows further
determinations of space that are hardly graspable by any other
means. Here we are confronted specifically with the problem,
touched upon above, of orientation in attuned space.
First, a question emerges with regard to the lived body and its
expressive movements. What are its characteristic movements in
attuned space and what can be inferred from them for the body's
orientation?
Most decidedly this is not a question of how the lived body as
one's own is experienced in attuned space. In being attuned I am one
with things and with space. My lived body is included in such an
attunement without any intention on my part toward it, or without
its condition being somehow present thetically to my awareness.
Such an awareness can emerge suddenly with qualitatively unique
sensations-data characterizable in accordance with "postura! situ-
ation," "depth," and intensity as typical modes of the inner giveness
of one's own body.t
0
Yet in attunement as such there is no positional
consciousness of one's own body. lt is reflection that first places my
body, with its modes and manner of intertwining with the world,
back into the sphere of judgment. Yet the lived body of attuned space
is offered to reflection in a unique manner: here it is phenomenally
unarticulated. Torso and limbs constitute a complete unity in the
attuned body. They separate only in an oriented space where the
limbs assume specific functions. They can specialize themselves to
the extent that the limbs assume an almost independent role from
the trunk, which, determined by the limits, is as it were only set into
10. A bodily "inside" with "postural situation" and "depth" presents the
body itself as spatial. For the relationship between lived body and physical
body, see p. 52 of the present work.
The Attuned Space 33
motion by them. In expressive movement, however, limbs and trunk
are mutually interactive. In this mutual attunement and harmony of
the individual dynamic forms, we experience charm and grace.
These appear in walking, striding, and dancing, as well as in resting
and standing still. E ven pointing and grasping in attuned space is not
simply the protrusion of the arm from the trunk, as is the case with
the conscious intent of getting something in the space of action;
rather, the limb remains completely caught up in the dynamics of the
whole body.
If we consider such an undifferentiated body in its modes of
motility, we notice salient differentiations from its purposive move-
ments in the space of action. Buytendijk rightly criticizes Spencer,
pointing out that by equating charm and simplicity of effort, he
completely misses the essence of expressive movement.u Its unique-
ness and richness is precisely in the unintended power, with its
fullness of possibilities undetermined by any rational point of view.
When Schiller characterizes those movements that "correspond toa
feeling" as charming, he expresses precisely the effortless, aim-free
movements consisting of "the beauty of form under the influence of
freedom. ''
12
What does the absence of any economy in expressive movement
mean for attuned space?
We take a couple of steps back in arder to see something better,
turn around, move to the side in arder to allow something to have an
all-sided impact. In the space of action "there are" indeed such
movements and, seen from the side of an organism, there is no
difference between them. Expressive and purposive movements do
not differ in terms of their course but are distinct in sense. In the
space of action, a step back is an "unnatural," compelled movement;
a step aside to avoid an obstacle or to make room, or any turning
around, are avoided if at all possible. In attuned space all these
movements are noncompulsory, unintended, and yet obvious. This
is not because in our attunement we are not oriented toward our
movements- even in the space of action we are not explicitly turned
toward the activity of the body, but to our aims and the means
available to attain them-rather, it is the space that constitutes itself
without compulsion, while the space of action is experienced as a
demand.
Expressive movement offers an answer concerning the orientation
11. F.J.J. Buytendijk (1), Section F.
12. F. Schiller, p. 254.
34 Lived Space
of this space. Attuned space has no specific orientation. This does
not mean that it is completely without orientation; such a space,
as space of a lived body, is impossible. Yet it lacks a qualitative
differentiation of values among orientations. The movements all
essentially occur with the same spontaneity, equally effortless and
obvious. Attuned space is atropic.
This state of affairs is supported by yet another phenomenon. If
one assumes that the endless multiplicity of movement orientations
of an oriented space remains related to the triad of the elemental
pairs of opposites-up-down, right-left, front-rear-then it appears
that orientation can be differentiated only to the extent that the pairs
of opposites are differentiated. Yet just these differentiations slip
away in attuned space. The differentiations mentioned are, as
differentiations in orientation, primarily functional; they are condi-
tioned by the functional articulation of the lived body, which lends
space an asymmetrical appearance, We shall speak of this while
discussing the space of action. In attuned space the body is "prior"
to such articulations. This can be seen in the case of powerful
expressive movement and posturerevealing a high degree of sym-
metry.13 At any rate, what has too often been repeated is also valid
here: the experiencing subject observed by us can hardly discover
the atrophy of pure attuned space since other kinds of ontologically
grasped possibilities of comportment toward the world are already
accessible to him. Y et the attuned space as such do es not contain
differentiations of orientation.
In attuned space these differences contain another differentiation
of value, i.e., valence of mood. The elemental pairs of opposites are
qualitatively distinct as bearers of expression and significance.
Hence the opposition up-down is strongly pervaded by an atmo-
spheric and specifically religious difference, constituting a mythical
residuum carefully preserved in all cultural development. In the
early Christian form of life there were also magical meanings of left
and right. Thus the left side meant the honored, the holy, yet also the
demonic and despised. It had to be protected by adornments.l
4
13. In addition there are the conspicuous ethnological discoveries that the
known religious rituals of prayer, devotion, and contemplation are symmet-
rical in posture. Yet let us also recall the bodily posture in sudden fear, in
abrupt astonishment, the gesture of supplication, the gesticulation in doubt
or in sadness, the exuberant joy, the posture of staring off into space, etc.
14. See J.J. Bachofen. In expressive experience there is generally a
noticeable residuum of myth, to the extent that the mythical structure of
The Attuned Space 35
The relationships prevailing between the mode of movement and
the determination of orientation are particularly clear in dance. As a
paradigm of expressive movement, dance reveals in a greater relief
all the moments that are given only weakly in attuned space, since
they are intertwined with aspects of movement belonging to oriented
space and thus cannot be grasped in their purity.1
5
consciousness can be considered as an early form of historical conscious-
ness. As E. Cassirer has attempted to show, the space of the primitives
possesses its own arder and articulation solely within physiognomic char-
acteristics. Certainly for him mythical space is merely a primitive precursor
of scientific space; the modern consciousness of space is specified by its
total liberation from such mythical elements. For him the expressive
experience of modern consciousness remains limited solely to the sphere of
I-Thou encounters, and thus is inappropriately narrowed.
It would require an entire investigation to deal with the problem of the
space of primitive peoples living today. These issues must be avoided
precisely because it is questionable whether the phenomenological-
descriptive method is adequate for them. The old demand of Schelling to
understand the mythical world not "allegorically" but "tautegorically" is
being reassessed by ethnology. The world of the primitives cannot be simply
delimited as primitive in comparison to ours; rather, it must be interpreted
as a self-contained whole with its own structural regularities. The magic and
myth must be understood from "the presuppositions characteristic of
peoples of these times" in arder to acquire a categorical system befitting the
differing consciousness-structure of these peoples. This is the requirement
proposed by A. Gehlen (2), p. 10, who contrasts it sharply to the method of
"understanding" which merely "starts with the present and moves toward
the past." Undoubtedly, this demand is applicable not only to ethnology but
to all historical sciences which take history seriously as history and as such
manifestly present phenomenological description in the broader sense with
a diversified field of work.
Nevertheless, we must not overlook the limitations of phenomenological
description in the specialized ethnological areas. Our contemporary con-
sciousness of the primitive mentality is notably an alien consciousness,
which always arises in the medium of its own categories. Even where we
employ methodological procedures that take seriously "the presuppositions
of such a mentality ," these eo ipso en ter the "understanding" of a conscious-
ness that requires, according to Gehlen, a structural change of its own. It is
only with such an explicitly self-critical limitation of its claims that an
ethnographically oriented "phenomenology of space" would be justified. At
any rate, phenomenology today would find a rich accumulation of ethno-
graphical materials, especially in the areas of the pictorial arts (see the works
of E.v. Sydow, H. Tischner, H. Khn, H. Read, H. Werner).
15. P. Valery interprets dance poetically and F.J.J. Buytendijk (3), pp.
36 Lived Space
Dance is a motile and complete oneness of torso and limbs, a
playful exuberance of dynamics and a beautiful aimlessness of
specific movements for which there is neither a point of departure
nor an aim, neither a beginning nor an end. It is an entirety of
movement. Just as it is indivisible into parts and pieces, so also its
space is not graspable in terms of a series of points on a path and a
multitude of locations. That dancing lacks any fixed and determined
orientation is quite obvious: while turning we are moving forward,
while moving backward we are stepping onward, while moving
onward we are returning-all this appears "impossible" in an
oriented space. The movement continuously assumes its space and,
so to speak, tenses it anew with each phase.
The previously underscored state of affairs ppears here from a
novel side. Each movement occurs not only in a spatial, but also in
a temporal-rhythmic, succession of movement phases. If we remind
ourselves of the role that movements have for attuned space, if we
keep in mind that attuned space is accomplished as a movement-
space, then we are once again offered a reason to speak of it as a
space-time.
6. Attuned Space as Space-Time
There are three states of affairs pointing to an interconnection
between space and time: attuned space as a form of executing
139-49, analyzes it psychologically. Its relationship to space was first
emphasized by E. Straus (1). Straus takes the movement of dance to be
expressive movement as such and attributes to it a particular _"presenta-
tional space." Since he sees an essential relationship between dance and
music, he concludes that it is through music that the structure of space is
first "created." For him, the "presentational space" is one of sound and thus
has a structure determined by sound. Straus's treatise is distinguished by its
refined observations and nuances and appropriately shows the general
interrelationship between space and movement. Yet it lacks sufficient
methodological strictness and results, on the one hand, in an unclear and
unrelated multitude of "spaces," and on the other, in a premature
absolutization. Thus an independent audial space is not guaranteed just
because the space of tone and noise ca-determines the structure of such a
space. That Straus could attribute to them a space-constituting function lies
in his inappropriate narrowing of the space of expressive movement to
"presentational space." Only subsequently will it become clear wherein lie
the fundamental conditions of space-constitution (see Part One, Section II,
of the present work).
The Attuned Space 37
movements, theii co-determination by something temporal (sound),
and their horizonal limitation through nearness and remotness as
spatio-temporal phenonena.
In arder to examine this interconnection more closely it is neces-
sary to touch briefly upon the problem of time. Yet it must be
explicitly emphasized that within the framework of our problem we
offer only a rough survey, a few hints and nota complete presenta-
Han.
The interconnection between space and time is usually limited to
the notion that space, taken as a location of points next to one
another, is related toa now, a temporal "point." Conversely, space
belongs to time insofar as the "flow" of time is represented as a
one-dimensional formation, as a straight line, i.e., as a spatial
continuum. Yet so conceived, the relationship between space and
time is only a loase proximity, a mere "also." Space appears "in" the
temporal point only as a postulate of conceptual completeness. It
reminds us that "next" to space there is also time, to be thought of in
the mode of now. In any case, the now allows everything spatial to
remain as it is. Space "is," while time "flows," and in each temporal
point space remains the same. Things change in it with time, but
space itself remains timeless.
M. Palgyi can be considered to be the first thinker who has gane
into the problem of the interconnection between space and time in a
persistent and original way.
16
With his conception of the "flowing"
space where each temporal point has a corresponding world space,
and each spatial point a corresponding temporalline, there appears
a new point of departure for subsequent space-time researches. It is
remarkable that it was not philosophy but physics that took posses-
sion of it!
1
Palgyi sa:w such an appropriation of his conception as a
crude misunderstanding and turned against it with indignation. Yet
in fact the alleged misinterpretation of his "flowing space" into
Minkowski's space-time continuum was accomplished so easily
only because in truth even in Palgyi the space-time interconnection
is seen as nothing other than a mere coordination. His space is "more
flowing" only because it is in the flow of the "adjuncted" time.
Nevertheless, there is a difference whether space is more flowing
because of time or whether it is flowing of its own accord and thus
conditions the flow of time itself. Palgyi fails to notice that his
chosen point of departure for this question is conceptually unfavor-
able. Like the physicists, he begins from the assumption that space
16. M. Palgyi, pp. 1-20.
38 Lived Space
and time are mathematical manifolds of points, that they confront
the thinking subject as objects and can be coordinated one with the
other after they have been grasped separately.
But what if we are dealing with a space that is not merely a
manifold of points, and that is not merely posited as the object of a
judgment, but is rather lived? What if time is originally not a
homogeneous, mathematically differentiated series of points, but a
lived time, a present that binds future and past? Would space and
time be merely coordinated, or is there an entirely different kind of
connection?
If our investigation of time were to correspond with our planned
investigation of space, then, in accordance with our method, we
would first observe how time is possessed by th subject----rnot as an
objective thesis, an object for consciousness, but "ekstatically" in
his lived comportment toward the world. From there we would
move to the analyses of objective time. It will become clear in
subsequent contexts why we are following the reverse procedure. At
first we shall discuss briefly both the question concerning the mode
of givenness of time in consciousness and that of the constitution of
this time consciousness.
Time is experienced as flowing: all events, all changes, and all
duration, occur "in" it. "In" it the subject knows himself and the
beginning and end of his corporeal existence. As such, time is
primarily given not as a change of the contents of consciousness but
as a happening of the world. Everything happens, occurs, runs its
course in it; in it there is enduring, abiding, beginning, and end;
Time is originally given as being-conscious of an event in the world.
A momentary occurrence is able to show the flow of time more
clearly than an enduring one. Experienced as now, in the mode of
the now, in the privileged givenness of originary vividness "in
person," the now is already past and has become something that has
been sinking continuously and irrevocably into the "depth." Finally,
it is extinguished in a completely empty background, inaccessible to
consciousness. Consciousness is aware of this vanishing on the basis
of its capacity to follow the now in its modifications, to trace it
retentionally, although not as far as one would wishP What is it
here that persists, what is it that in a specific retentional phase does
17. See E. Husserl (4), # 77, for the distinction between retention and
reproduction. Subsequently we shall deal with Husserl's hyletic data.
Protention, for which there are essential analogues, must be left aside in this
brief sketch of the time problem.
The Attuned Space 39
not allow the now to change to another now but holds it as just
having been and thus as the having be en of a now? This persistence
in the stream of time would be incomprehensible if the simplest
world event did not have a specific sense-content, if the conscious-
ness of the now were not a being-conscious of a sense-bearing now,
understandable through all the phases of retentional changes. If the
now of the sound of a bell were to become the whistle of a
locomotive, and thus to be experienced differently in each point of
the retentional continuum, then it would not be something that has
been-it would not be the retentional modification of a now.
The identity of the sense guarantees the relationship toa now of
something that is passing and makes comprehensible the "now" and
the "having been." Furthermore, the identity of the sense-content
implies that the given time has a holistic structure even if it is
distended across the ekstatically lived time. Although distinguish-
able in terms of past, present, and future, this triplicity of temporal
modes is not separable; the just-passed and the just-coming are
co-determined by the now, and it by them. Moreover, the phenom-
enal now is not a discrete point; taken chronometrically, it can
persist. Further, there is another remarkable characteristic that
distinguishes the originally given time of consciousness from chron-
ometric time.
The previously observed singular event as something given is an
abstraction. Were we to follow a number of successive events in
retention, the su'ccession itself would also be obtained in retention,
and thus the originary time of consciousness would assume an
unequivocally directional determination. The latter event cannot
surpass the earlier (which can be the case, for example, in a
genuinely false reproduction). Earlier and later are retentionally
irreversible; time in its flow never reverses its direction. Neverthe-
less, outside the narrow sense of retentional modifications, its
temporal distance undergoes a specific change. With progressive
obscuration, the distance collapses toward an "infinitely remate"
nebulous point where all distinguishable contents vanish. The
originally given time is oriented, finite, and perspectiva!, as is
oriented space.
But does not the retentional continuum continuously lose its
orientation precisely because its point of orientation, the living now,
is itself in "flux"? Does it not require in turn an orientation toward
something that "holds" this continuum? This something as such
must have extremely contradictory characteristics: on the one hand,
it must necessarily be changeless, fixed, and identical; and on the
40 Lived Space
other, it must not remain outside of the temporal flow. Here appears
the riddle of what we call consciousness in the form of self-
consciousness. This consciousness does not vanish temporally with
the changing objects but maintains itself as the selfsame. 1t outlasts
all changes and retains its own identity in the flow of time. Yet it is
this very consciousness that can detect its own identity only in the
flux. This does not mean that self-reflection arises along with the
being-conscious of time, such that consciousness thereby gains
access to itself, but only that the conditions for the possibility of a
time-consciousness as such are constituted in it. Time can thus be
given objectively only to a being who knows itself.
Reproduction must be distinguished from retention. Reproduc-
tion, as a recollection, is the way that natural consciousness moves
as time-experiencing. While retention is a flowing after along what
was once present, reproduction is a representation of what has been.
lts self-orientation to the past is a spontaneous accomplishment of
consciousness. Consciousness is capable of "transposing" itself to a
locus or a duration of the past that is illuminated anew as if it were
in the modality of the now. Since we live "in" remembering, this
new now is neither the originally given now, present "in person" in
the lived experience, nor is it present as existent, since what is "in
person" is the now of a past. Meanwhile, this now has long since
been surrendered to the stream of the retainable; it is repossessed in
reproduction and through it "made" into a living present. Yet here
the retentional perspective undergoes a characteristic transforma-
tion.
Since recollection is phenomenally a fulfillment of the retentional
continuum, whose temporal stretches are seen perspectivally, it
seems that reproduction cannot give us any absolute size constancy.
Yet with representation as a spontaneous act of self-transposition
there appears a new problem. Taking the fact that in the act of
representation the newly attained now d o e ~ not have the mode of
givenness of originality, "perspectivally" then means: perspectiv-
ally to any re-presented now, i.e., in principie to any arbitrary point
of retentional continuum. Yet each duration in proximity to the now,
the "freshly" reproduced, points to a slight perspectiva! shift. If it
were possible to fulfill each position of the retentional continuum
with a "new" (re-presented) now, then the perspective would
thereby be removed, the "actual" duration of the individual tempo-
ral stretches would be regained in (re)experiencing, and time would
be homogeneous. lt would be necessary that any point whatsoever of
the retentional continuum be reproducible, whic;h in principie is the
The Attuned Space 41
case; yet, in addition, each of these positions would have to be
reproducible "now," which in essence is not attainable. Instead, it is
possible to repeat the recollection of a reproducible event an
arbitrary number of times; the reproduction can turn back at will to
the same event of the pastas often as it wishes. In such a repeatabil-
ity there is a corrective factor for the perspective of time. Since two
recollections are never the same, in that they are enacted from a
different now and are continuously motivated by diverse factors,
such repetition of reproduction offers the possibility of freeing an
event of the past from its perspectivity. Moreover, when an event is
reproduced, each of its recollected now-points can be rendered into
a new recollection and this in turn can be repeated. This means that
the structure of reproducible time is similar to that of homogeneous
time. Regarded phenomenologically, the free motility of the repro-
ductive glance prepares the ground for the construction of chrono-
metric time.
Despite the many simplifications of the states of affairs, the
discussion presented above should not deceive us concerning the
real complexity of the temporal problem. Essentially it had to do
with the modes of givenness of time in natural consciousness. It
remains in a sphere of temporal givenness that is not the most
original.
Time offers itself to reflective analyses in a mode in which it is not
"given," or consciously presentas flowing, but in a way in which it
is appropriated and lived, in the primordial sense of the word,
without thetic awareness of it. Heidegger developed this "ekstatic"
mode of time appropriation from the care structure of Dasein.ts Time
in this conception of an enraptured ekstasis is even farther removed
from the time of consciousness as a continuous, homogeneous series
of the one after the other. Unlike its appearance in objective
consciousness, where time is constantly related to the present in its
three modalities, time here is a unity of the three phases of
"ekstases." Here the future is not later and the past not earlier than
the present; rather "Dasein is temporality as past presencing future";
it is constantly "contemporaneous." As will become obvious, this
contemporaneous structure of lived time can be grasped only as the
"time of Dasein," but not as an inner-worldly time. It has its
ontological ground in the temporality of Dasein as the sense of its
being.
Undoubtedly Heidegger was able to articulate structural charac-
18. M. Heidegger, 65-71.
42 Lived Space
teristics of lived time that had to escape Husserl because of the
latter's orientation toward the constitution of objectively experi-
enced time in pure consciousness. But how is the temporality of
Dasein as contemporaneiety to be understood in relationship to the
inner-worldly time? What the latter means for the being of the
subject, and what it is in its own right, was never interrogated by
Heidegger. A more exhaustive investigation of this question would,
however, lead us too far afield.
After this digression into the question of the problem of time
we renew the discussion of the interconnection between space and
time.
If Palgyi's proposal turned out to be inappropriate for showing
structural unity between space and time, the analysis of time-
consciousness could only show that the "time of intuition" is
complete! y free from spatial moments. Its relationship to the space of
intuition was only one of analogy, implying a deeper layer of
interconnection that was not yet itself revealed. In arder to discover
a plausible interconnection between space and time, showing not
only that they have a relationship of coordination but that temporal
components are traceable in the spatial structure and spatial com-
ponents are traceable in the temporal, the interconnection must be
sought in the forms of lived spatiality and temporality. And these
cannot be grasped objectively but must be presented unthetically,
not known but lived or accomplished.
That such a search for space-time unity did not arise only from a
speculative need was evidenced by the phenomena in attuned space.
What leads further into the question raised above is precisely its
characteristic form of accomplishment in expressive movement. If
one seeks the mode and manner in which not only spa:ce but also
time is obtained "ekstatically," one finds it precisely in living
movement. While this is true of both expressive and practica!
movement, there is, nevertheless, a noticeable difference between
them. While the latter realizes the "ekstatic" unity of the three
phases of time in its accomplishment, in objectifying reflection it
allows this unity to be incorporated into objective time. The former
is completely incapable of such an incorporation. Even reflection
upon the expressive movement does not succeed in grasping it as
"occurring in" a specific time. Expressive movement is nota process
that begins and ceases, commences "now" and breaks off "then." It
is upsurging and resounding without fixed limits; it has no disrup-
tions in objective time. Even its stillness is an arrest in the whole of
movement, which not only contains or includes past and future in
The Attuned Space 43
the present, but is pure presencing, pure contemporanaiety. Subjec-
tively and objectively, expressive movement is the paradigm of an
"ekstatic" temporal wholeness to be thought prior to any differenti-
ation into temporal modes.
Any state of affairs must be grasped from its time characteristics in
such a way that attuned space, as a specific space of movement, is
constituted through something temporal. Temporality does not
mean here a process in an already present objectifiable time, but a
grounding of time, time as "ekstatic," projecting. Corporeal move-
ment can be formulated in total indifference to space-time. At the
same time, it is corporeal movement that first of all incorporates time
into space and the latter into the former. This is a highly inadequate
way of spmiking, stemming from a thought that is dominated by two
separate "forms of intuition." Yet in corporeal movement there
seems to be an ontological ground for an originary unity of space-
time (pp. 145ff.).
It is quite remarkable that the traces of time in space show up
much more clearly, phenomenally speaking, when the subject lives
in it more "timelessly." In attuned comportment toward the world
there is no time for the living subject; a being who is essentially only
an attuned corporeal being knows nothing of temporal flow. Time is
dissolved in the experience of attuned space--and in the reflective
analyses of this space, time will be grasped as a moment of space. It
is distance in particular that allows the cognition of the temporal
moment of space. As spatio-temporal phenomenon, distance
determines the motility of the horizons; as spatio-temporal, it is the
limitation of the "metaphorical" spaces. Distance delimits the
attuned space both as temporal space and as attuned time-space.
lt is on this account that in the nearness and remoteness of at-
tuned space, spatial and temporal determinations are mutually
pervasive.
7. Attuned S pace and the Experiencing Subject
The structure of attuned space appeared in our investigation of its
fullness and emptiness, of place and situation, of its nearness and
remoteness. Its essential characteristics presented themselves in the
reflective analyses of the experience of space.
At this point we can offer two valid objections. Philosophical
consciousness since Husserl has taken care to trace the distinctions
between something and the experience, perception, and cognition of
that something. We must not overlook that in our special problem we
44' Lived Space
are not concerned with noetic-noematic unity, insofar as attuned
experience is not an intentional orientation toward something. Still
the question could be raised concerning the relationship between the
experience of space and space itself.
Is space itself inaccessible to the analysis of the experience of
space? The answer to this question was already prepared in the
considerations of space and expressive movement; it requires com-
pletion and deepening.
lt appears that experience is here to be understood only with a
grain of salt. Attuned space does not confront the experiencing
subject as something independent, a being in itself that must first
opera te in arder that one may "react to it." S pace does not ha ve an
existence of its own, separated from the subject, to which the
subject should establish a relationship; as a space of my movement,
it is much more space through me than my experience is through it.
The strictly reciproca! implication prevailing here between space
and the experience of space can more easily be shown through its
characterization as an event than it can be subsumed under fixed
concepts. What can be grasped in immediate perception as an
encounter between subject and space appears all too easily to be a
paradox.
Attuned space offers itself in its fullness.lt appears that the fullness
is nota mere methodological expedient, a specific approach to space
that could be grasped on another occasion by other means; rather,
space is this fullness itself. Its vanishing is thus not the disappearance
of something in it but a loss of something as a whole. The phenom-
enon of disappearance best reveals the reciproca! relationship be-
tween space and the experience of space. While attuned space, with
its full physiognomic content, corresponds on the experiential side
to the uninterrupted fullness of psychological impulses announcing
themselves in the richness of expressive movements, the obliteration
of attuned space is only one si de of the reciproca! relationship. On the
other side, the loss of attuned experience extends all the way to the
"emptiness of heart" of which Scheler spoke and for whom it is the
"originary datum of all concepts of emptiness as such."
19
The indif-
19. M. Scheler (3), p. 298. In addition, see H. Tellenbach's report concern-
ing the spatiality of melancholy. According to him there is frequently found
with these patients a disturbance of the relationship to the space of action,
which Tellenbach, following Heidegger's Dasein-analysis, suggests is a loss
of "nearness" in the sense of making room for equipment (p. 292). However,
the "emptiness" of melancholy does not correspond in a phenomenologi-
The Attuned Space 45
ferent or callous person fails to notice anything that addresses him;
to the extent that he feels ernpty himself, he stares "into ernptiness."
With the phenomenon of absolute emptiness as well as with that of
concrete fullness, the reciproca! relationship between space and the
experience of space is visible, a relationship that can hardly be
thought in sufficient intimacy.
Unlike things, attuned space is not "outside" me. It "surrounds"
me, it is about me-this is its mode of givenness. But I am not in it
in the same way as things are in it. Through my experience, I am
spatial on the basis of my possibility of being an experiencing being
that is an expressive, motile, living being. Attuned space is with me
as an accomplishment of my attuned being, relating to it in mutual
conditioning and fulfillment-this is its mode of being. In this sense
its being exhausts itself in being a being for an experiencing subject
and above this it is nothing "in itself."
An immediate objection arises: attuned space must be "merely
subjective." The concept of subjectivity, as well as its correlate, is
engulfed in a multitude of meanings. Even if such concepts, are
gnosiological and not ontological, the purpose of our investigation
requires their brief discussion.
For one, subjective means that which is appropriate to the ego;
subjective is everything that is in me. This relationship is
determined and limited through the (unreflectively experienced)
relationship of my ego to my body. What is decisive here is not the
body itself but this relationship. Bodily possessioJ).s-limbs,
organs-are nothing subjective. Subjectivity requires a relationship
to an ego that is distinguished from its lived body even by a
non-reflective consciousness. That this relationship of the body
signifies essentially a subjectivity means only that the ego grasps
itself as an ego of a lived body. It is not that it is a lived body as
such, but that it is a lived body of an ego, of a self, that makes it
capable of distinguishing what, in an ordinary sense, is one's own
and what is alien. Subsequently it will be shown that the sense of
the relationships "in me" and "outside of me" is based on this
relationality of lived body to ego (See pp. 140 ff.). In this sense my
perceptions, surmisings, and stirrings of feelings are subjective; they
belong to me, they are my "own." The correlate to this subjective
cally precise way with the ernptiness we are refering to. Tellenbach's
observation that the inner ernptiness "corresponds" to the ernptiness of the
world and the ernptied space "intrudes into the inner ernptiness" (p. 16), is
nevertheless worthy of notice.
46 Lived Space
side is what is alien to ego, alien to me. Things and their
relationships do not belong to me; even "alien" persons are in this
sense not subjective.
Only that which is m y own can be "merely subjective." By this we
mean deceptions, errors that appear as such in the disruption of a
coherent understanding, in the cancellation of an experiential con-
text, of a continuity of sense. They can be discovered by me or by
others. Because of the possibility in principie of such a discovery,
there appears a further meaning of the subjective. What is subjective
is that which an ego calls its own, what is with me and m y ego. Thus
in a primary sense the other, the alien, ego is subjective; it is a
"subject." l':go-ownness is then not only my own but an ownness of
each ego, all egos. lts opposite concept is all that is aliento each ego,
the totality of the "objects."
This subjectivity is of a singular kind. Each ego is one by virtue of
its lived body; the Hved body is a mode of givenness of "my" ego as
well as of "another" ego. Another ego and my own ego mean
corporeal ego and nothing else. But as such, ego has surpassed its
own, and I have surpassed my own, corporeity. This transcendence
of the body grounds the subjectivity of the subject as universal. lts
correlative concept is objectivity as intersubjectivity. Language, art,
history, and science are objective in this sense.
What sense can then be attributed to the claim that attuned space
is subjective? lt cannot be subjective in the first sense; it is not "in"
me but "about" me. The serenity of a landscape is not that of my
sensations but is something in the landscape. I can experience it in
contras't to my own (ego-own) subjective sensations. In this sense
. attuned space is also objective.
Obviously this claim does not mean that attuned space jg objective
in the sense of intersubjectivity; it is not a homogeneous space that
is the same for all subjects. lt must be admitted that in its expressive
fullness the attuned space, with all of its specificity and uniqueness,
appears "always" as my own. Yet I know myself in it at the same
time with others-or without others-alone. Solitude is comprehen-
sible only as an absence of others. To become aware of solitude is in
its own way to co-experience the other. Thus attuned space is
comprehensible only as a possible space for the others. lt retains an
intersubjective moment and therefore an objectivity in the latter
sense indicated. lts relationship to an experiencing subject, which is
certainly its general characteristic, cannot be confused with subjec-
tivity in the sense described above.
What can appear as doubtful conc3rning the existence of at-
The Attuned Space 47
tuned space is, in any case, the fact that in experience it cannot be
grasped objectively. This is not due to attuned space, but to the
subject, which, in the mode of being of attuned experience, relates
itself sensibly to a world without standing over against it as a
"sub j ect."
J 1 {l \Q"'--\, !(
1
'v , (:J
1
\ '('
Chapter Two
The Space Of Action
1. Preliminary Remarks
Lived space is not exhausted in being solely an attuned space. This
does not yet encompass the totality f spatiality, which, as experi-
enced, has its specific characteristic of being related to a corporeal
subject. Attuned space turned out to be a space of expressive
movement. As such, it has the uniql.ieness of being free from
differentiations of orientation. This is its profound difference from
< both of the other forms of experienced spatiality.
The concept of orientation is characterized by two factors. Orien-
tation presupposes differentiatable zones, determinable loci, posi-
tions, a here and a there; orientation is always an orientation
from .... toward. Furthermore, it includes the possibility of move-
ment appearing as "directed" and oriented. Expressive movement
--l. has demonstrated that not all movement is oriented. It has also turned
out that this is closely associated with there being no point of refer-
ence in its space. Thus lived space can be oriented only to the extent
1/11\f,-that there is a formation of a center in it. Orientation and centering
of space are only two different terms for the same state of affairs.
If there is to be a form of lived spatiality manifesting a univocally
determined orientation, then the corporeal subject must exist in it as
a lived body that can be grasped univocally as located here in
distinction to each there. In this way the corporeal subject appears in
two respects: as acting body it is the point of departure of goal-
oriented activity; as unity of the senses it is the'point of reference of
sensory intuition. In accordance with these two modes of comport-
ment there appears a distinction between the space of action and the
space of intuition.
The primary formal determination of the space of action is the
"wherein" of possible activities. The concept of activity will be
understood as the realization of a project through the lived body and
48
The Space Of Action
1
49
.-/
its members. Thus we avoid the that the lived body is an /
implement, a means in order to .. : . This frequently employed
instrumental image is factually incorrect; the lived body is funda-
mentally distinct from all i:hstrtiments. It stands in an irreversible .,
relationship them by having and manipulating. ThiSISiOtThe
into the problems indicated. Yet at the outset the acting
body must be seen as capable of manipulating implements. Subse-
quent observations will consider the form of bodily activity employ-
ing certain tools as means. But is the way in which bodily activity is
understood of signifigance for its space of action? This question
introduces into our investigation a specific prejudice concerning the
relationship between the space of action and the acting subject. It is
the task of the subsequent analysis to take a stance toward this
question and to show what it means to say that all activities occur
"in space." It cannot be decided at the outset whether the en tire
structure of this space is indifferent to the fact that its center contains
a being who employs implements and instruments. .
But the subject does not merely employ the implements, which
leads us into a problem. The subject of this space understands the
use of equipment in whose manufacture the mathematical construc-
tion and the laws of exact science play a role. These instruments,
strictly speaking, "apparatus" presuppose the geometry of measure
and the theory of physics. It may be objected that this unexpected
intrusion of the natural sciences into the corporeal sphere of the
subject disrupts the basic methodological principie of this investi-
gation. This principie requires that the mode of comportment of the
lived body be the sale point of departure of our investigation.
Nevertheless, it cannot be denied that the observed subject's behav-
ior does not differ phenomenally in the least when he handles a fS
constructed apparatus instead of a simple implement: in the actual
process of activity, both are equally means toward the attainment of
something. Otherwise he would not be an acting being oriented
toward the utilizability of things.
At first glance this may suggest that technical construction need
not be considered in the space of action. The silent assumption
would be that nothing is changed through its structure. This assump-
tion has been always made whenever the space of action entered the (
field of philosophical interests. Credit is due to existential philoso-
phy for having pointed out this assumption within the framework of
its problems. Yet its conception of the space of action does not
recognize an apparatus as a problem. Of course no methodological
inconsistency lies therein; the existential concept of being, i.e.,
50 Lived Space
"Dasein" (as Heidegger's "care" and Merleau-Ponty's "Etre engage")
and the limitation to the sphere of everydayness is sufficient for the
existential point of departure.
lt is different with us. Indeed, we retain our point of departure; we
observe the subject in the space of action, in which he comports
himself and understands himself in an unreflective attitude. Phe-
nomenological completeness requires that we consider his manipu-
lation of apparatus; this is no less a phenomenon than pre- and
extra-scientific praxis. That the active subject engrossed in the world
of work may have no insight into the specific mode of mediation
between the lived body and constructed apparatus does not justify
our discrediting it in philosophical investigation. Yet we face a
specific problem: the manipulation of apparatus suggests a state of
affairs whose signifigance can be understood only after the investi-
gation of geometry, i.e., its so-called application to lived space. The
appearance of fact at the head of our investigative progress intro-
duces factors of technology the comprehension of which assumes
that the aim of our progress has been attained. Moreover, the
clarification of such factors assumes that the investigation of the
lived spatiality as a whole has been completed.
This phenomenologically most disquieting situation places us
befare two alternativas: either we must curtail discussion of those
phenomena that can be sufficiently articulated in the attained level
of investigation or, while including all the phenomenal contents, we
must make anticipations of areas to be analyzed subsequently and
declare such anticipations explicitly through reflection. Previously
discussed reasons suggest the appropriateness of the latter course.
The access to attuned space turned out to be its fullness. The
expressive characters of its things constituted its atmosphere. The
expressive world dissipates in the space of action. Here the expres-
sive characteristics of things vanish into the qualities required for
their utility. Thus they lose their effective and communicative
physiognomy; now they reveal their suitability or resistance "in
view of " a goal. This is not to say that the glance now penetrates a
kind of expressive level of an entity and reaches its genuine
determinations, its "in-itself." Expressive characteristics and prac-
tical properties of an entity do not relate to one another as a surface
to a kernel. An entity is, in its mode of being, phenomenally whole
and complete in its determinations; it has them as a whole. Further-
more, they do not comprise its different "sides." This way of
j speaking obscures the fact that we are confrontad by determinations
of two distinct regions of sense, constituting themselves only in the
The Space Of Action 51
subject's orientation and perspective. The applicability and the
usefulness of entities are first given as such in a particular grasp;
they are opened in projects of activity, and apart from these projects
they remain incomprehensible. Furthermore, this is not to be under-
stood as if an entity were apprehended thematically as the being-
there of Dasein or the coming-before-us of things. Active engagement
with things has a specific mode of seeing, appropriately character- . . .. \
ized by Heidegger as "circumspection." Its discovery rests in the e 'f(' ~ \ ~ ' '
implement, whose mode of being is ready-to-hand.
20
The structural analysis of implements in the space of action,
nevertheless, runs up against sorne difficulties. This space is
ontologically relative to a temporary project, to a specific situation of
the acting subject. Thus it contains a temporal moment; the space of
action has a dynamic texture. Is it possible to discern anything about o
it? In its constant transformations, must it not continually escape
conceptual boundaries? It would be so if each project were an
absolute beginning, if each actual space were completely distinct,
and if between them there were no constant transitions and perva-
sive regularities. Asan acting being, however, the subject appears in
his historicity. He finds himself in an already formed world of work
that is not his own creation. Immersed in it, he also participates in its
constitution. While transmitting and at the same time shaping what
he has acquired, he realizes a relatively enduring project-such as in
his vocational decision-from which the actual attains sense and
meaning. The spaces of action of greater historical dimensions,
formed in common cultural labor, comprise at the same time the
framework for momentary individual projects and their "spaces"
which are neither rigid nor motionless but can be supported,
extended, or negated in actual activity.
It is not necessary to follow this problem any further; rather, we
must ask: how does a space of a historical being present itself, a
20. M. Heidegger, 15-16. We appropriate these differentiations without
following Heidegger's ontologicaL conceptions. Specifically, we do not
accept the validity of th![ bntol,o,gi_cal j)rirl!Y>of circumspection, in contrast
to Heidegger's notion of the"free comprehension" of pure sight The claim
that circumspection is more primordial and that sight is founded on it has
for us no sufficient phenomenological grounds. Heidegger overlooks the
fact that circumspection already implies sight, and that the latter is
co-constitutive of circumspection. Moreover, the comprehension of some-
thing as present at hand must place in brackets its qualities of being
ready-to-hand.
----
52 Lived Space
?
space that has in its content a being who is no immersed in

attunement but is oriented toward the world and strives toward
aims?
2. Place and Region. The Space of Action as a Topological
Manifold
Like attuned space, the space of action is not justa mere multitude
of points in three dimensions. Yet structurally the latter does not
conform to the former. It will be shown that the texture of the space
of action has become looser, that the role played by the part in the
whole has become different, less far reaching.
The space of action is articulated according to places and regions.
Place is thelocus of what is usable, discovered by the acting body. In
essence: H1ese delimiti:ions agree with Heidegger: place is the lo.cus
where implements belong. (Included are the privative modes of not
belonging there, lacking, being in the way.) In belonging thf)re, tl}e
thing possesses a relationship to "its" place. The same
appeared in it must be understood
differently. In attuned space the place. belongs constitutively to an
entity not only in the specific mode of being as a carrier of
-:: expression, but as an essential co-determinant of, space as a whole.
In the space of action this is true only in a very restricted sense.
That something ready-to-hand has "its" place is determined prima-
rily by a moment of its duration: it is found there "customarily,"
most of the time," and it has its "usual" place there in accordance

"t... S
with the requirements of the acting subject. Yet its belonging to a
place is not identical with its appertaining to a place and is not a
constituent of its utility as such. Its place is variable within a broad
limit, without a loss of its character as a specific instrument. Within
limits to be specified more closely, of the ready-to-haJJ.d
things can in with()l!t t}e lqss of!he mocie()f
being ofthe-ready-to-h'u.d. -This variability is rather a constituent of
the place of an instrument and opens the possibility for "dealing"
with things.
This comprises a fundamental difference from things as bearers of
expression. One may compare paintings in a gallery, arranged for
"proper viewing," with the same paintings in a workshop, merely
lying "to hand" for framing. Things that are bearers of expressive
a value receive fewer and less carefully selected places when they
become mere objects to be used for a purpose! Of course, the place of
the ready-to-hand is not sorne arbitrary location. In its mode as being
The Space Of Action
53
I
J)(ir v' {:lt\jr,
ready-to-hand it is relative to the subject's mode of being asan acting
subject and thus to a great extent independent from place; yet as
in an thuSit must
be handy.
Handiness can have two meanings. On the one hand, it means an
adaptability to the organization of the body. That which is handy in
this sense is what is "tailored to the body," such as hand tools and
instruments of daily practice. It is a handiness of means for the
realization of projects. On the other, handiness means "having in
hand" something light and comfortable, reaching for the useful in a
shorter way, with lesser hindrance, etc. It contains a specific
principie of economy that will be considered subsequently. Both
meanings must be separated: a thing that is handy in the first sense

\cv\ (
can be unhandy in the second sense, as being "there" at this _ _
moment. The handiness in the second sense is a function of place. ( ov..N) IC\
This of places; it is decisive for the \I'PI'
space-of action as a manifold of - ---- -- - - - --- - ,
For a thing to llave "its,; place ils-an instrument means that the
place is constitutive not simply for its mode of being ready-to-hand
per se, but for its handiness within a project. In order to be handy, to
refer to a profect, it must be by the choice of a subject.
Yet its place is notan arbitrary the ready-to-hand .
as such can in principie be anywhere, even if its whereabouts is not
an arbitrary place in a system of locations. Heidegger mentions that
the place of an implement is discovered when it is missing. In such
an "absence" the implement simply "comes before us" as it vanishes
from being ready-at-hand. Nevertheless, it is the place that becomes
obtrusive as having something missing, even if what is conspicuous
by its absence is something ready-to-hand. "Something missing" can
only mean a missing implement, the search for which is precisely a
search for it in its ready-to-handedness - a search motivated by its
having this mode of being as such.
21
The phenomenon of the search
reveals the "being somewhere else" of what should be ready-to-
hand. Even while missing it remains ready-to-hand with the possi-
bility of place variations; it will be missed precisely because its
handiness has assigned a specific place to it as its "there." It is
21. Heidegger does not distinguish between ready-to-handedness as a
mode of being of an implement and handiness as a characteristic of an '1 \.,
implement in an actual project;-thus the phenomenologically unjusti-(.---- e \
fied transition from ready-to-handedness to present-at-handedness in the \
1
_ \
"search." )''V<lO} tC!f
. \.A
54 Lived Space
prepared, placed appropriately, accommodated, misplaced-all
these characteristics which are not given for things in the attuned
space include a certain Qrovisionality of place.
An important note must be added. The place of an implement
. J,; 1 prescribed and determined by handiness in an actual project is
rc'""'!!'c '1 variable within specific limits, i.e., it cannot be precisely fixed. This
. \-1 of the region.
ot Until now the places of the ready-to-hand were seen in isolation as
individual there and yonder. The abstraction inherent in this view
must be revoked. In a project of action the individual implement is
always transgressed toward something further; it obtains its appli-
cability from a and in turn points to a
possible totality of involvement. The place of an implement in its
"there" is determined by the range of other places. Its "there" points
toa specific surrounding of other "theres." Its "where" is at the same
time "whence" and "whither."
and whither is the whole field of action; it is
__ It is possible to extricate relatively
independent areas from it, regions as entire places of relatively
closed connections. Each space of action is divisible into such
regwns. To demonstrate this more strictly would call for a more
exact analysis of the notion of a project than can be offered here. We
limit ourselves to thesuggestiorithafan activity presents itself in
terms of its stages of fulfillment, determined in accordance with
partial aims. This has todo with the manipulation and use of things,
building relatively independent and limited functional contexts
within a totality of instrumentalities of the entire project. Within the
entire space of action, these factors constitute relatively closed
1
space-manifolds, comprising what we have called regions. Their
relationship to the place of an individual thing leads to new and
important determinations that make the structure of the space of
action transparent.
The place of the ready-to-hand is determined through its regions,
although it is not precisely determined. It is not a punctiformal
1\) "where," but a somewhere within the limits of its region. The region
is the "leeway" for the free variability of place within which it
remains changeable in a restricted sense: it remains the same only
when it remains within the region. The ready-to-hand is allowed
certain displacements without its place ceasing to be the designated
one. Two things find themselves "in the same place" only when they
are found in the same region. Region is definable in this manner: a
place-manifold whose specification adequately satisfies the ques-
Cllv.A
,(1
1\W\N\
The Space Of Action 55 ,'e NL
tion concerning the whereabouts of an implement. "At the work
place," "in the desk," are such specifications, which are topograph-
ically exact as to the region of an implement in the space of action.
In principie, this includes the possibility for further specification
of any region. The extent of each region is relative to a project and to
the possibilities of activity Regions as such
natfrst eStaolisrum-ana opened but rather arise with what is
encountered in them. What is encountered determines the extent
and the limitation of their further articulation and
the possibility of j)f and thereby the
structuratwn of space. For a wanderer a sea 1s a smgle homogeneous
region confronting him as impassable, while the fisherman, the
swimmer, and the seafarer, in their differently motivated actions,
know how to discern its various regions; for them it is structured
differently. In addition, each individual project allows in principie
the nesting of regions through progressive articulation. Of course,
this articulation cannot be extended arbitrarily. The articulation
determines the degree of structuration. In this regard, different cases
must be distinguished.
In a fully structured space any region can be neste_d at will. We can
characterize such space by the fact that the nesting df regions can
continue in it without restriction. A thoroughlyiliuctured space
would be a space where each arbitrary sequence of nesting of regions
could converge toward a determnate place, and, conversely, where
every place could be reached through at least one such sequence of
nestings. In the thoroughly structured space, the process of nesting
determines place univocally as a point of this space.
22
Due to its
mode of establishment, it, too, is a region, although the smallest
region of this space. This do es not coincide with the definition of the
region as the "leeway" for the free variability of place. Yet it must be
that _a titQ!_D!lghly an
limit of the space of action. As a whole, it can never be completely
structured, but or less structured. We are talking about a
structured space that is neither completely structured nor totally
unstructured. This implies that it can have completely structured
parts and also totally unstructured partial spaces. The latter pre-
cludes distinguishable places and hence does sequence
22. As an extreme example one could think of a surface covered by tile.
Each tile in it has its place from which no shifting is possible. Each such
place is a "point" of the tile-surface, i.e., in its space there is no "there" to
be established with greater precision.
'V\()1
(/1
56 Lived Space
of nestings of regions. What is essentially unattainable is that the
space of action as a whole be completely structured or completely
unstructured. The latter is excluded due to its referential relation-
1 ship to the acting subject. As a bodily being in this space, he
understands his locus from other loci, his here from distinguishable
theres and yonders. This differentiability determines the orientation
of space. This means that the spce of -t!on, asoriented space, must
__ . - - -
It must be mentioned why the space of action is not completely
structured-more precisely, why at least one partial space must be
given in which the nestings do not converge toward fixed places.
Why must it have places that are not "firmly" established but, on
the contrary, can be determined only up to a space with leeway?
The basis for this was already suggested: the ready-to-hand attains
its place as required by our dealings with it. These dealings not only
allow but demand the inexactness of the locus given iritlie space o'
the of leeway is nota
topographical inexactude to by progressive activity;
rather, it is a positive determination accruing to the place of an
, , instrument. If leeway were lacking the things would be too dense;
'0J .. "'": (1 the acting subject would not be in full possession of his possibilities
8-f'r \ of action. Rather, he would object to the "overfullness"-without


\ the leeway of its place, the ready-to-hand would be completely
[L
0
unhandy! The variability of place within a region is constitutive of
the locus of the ready-to-hand. Place is thus topographically
graspable "only" up to a region-yet in the smallest region, the
space of leeway, place, can always be established with sufficient
exactness.
In any case, the leeway as such is not given thetically to conscious-
ness in activity. lt is co-posited circumspectively in the management
of so:r;nething without there being any intention expressly oriented
toward it. The acting subject is not primarily attending to a region,
,' (.vl of his project that
_the Only the encroachment of the too-dense
fl!\ revea1s the lack of leeway. Conceptually, this involves a view which
l\-.'-\\\

entirely different It is.only the intuitive


o\cL\'1 {v; \ v1ew that preserves d1fferentlatwns of place allowmg us to speak of
_ . place in activity and circumspection. In pure contem-
V (l V -
1
d" plation space appears with another structure. However, pure percep-
tion disrupts activity; the usable thing of our dynamic dealings is
rigidified into a static, isolated object. Robbed of its utility charac-
teristics, it is torn from its functional context. The project is annulled
The Space Of Action
?
57
o
and the space of action collapses. The pure space of intuition comes
into view.
The space of action turned out to be a manifold of regions where
possible nestings constantly converge toward one region. The small-
est region, as "leeway," determines the place of the implement;
place is univocaly determinable only up toa region. The constitutive
inexactness of place requires the region to be the smallest topograph- J
1
ically exact element of the space of action. What significance does V
this state of affairs have for the understanding of space?
This question is related to the final intent of our investigation
which is concerned with the phenomenological origin of geometry.
Meanwhile, taking the concept in a sense yet to be
discussed more precisely, we the space of intuition
as the source of geometry. The previous analyses demonstrated that v
the source must be sought much earlier, namely in the space of
action. Indeed, the space of measurement has not yet emerged. Y et ,
geometry presupposes a mathematical discipline-one which, al-
though exact in a scientific sense, still remains on the side of
all determinations of measure, which functions without point-
geometry, and whose basic element is the region: topology. Staying
with the metaphor selected, it could be said that the source of
topology is to be sought in the space of action ..
3. The Locus of the Subject in the S pace of Action
As with other things, the subject also assumes a place in the space
of action. Yet he has space differently than they. Things of utility
have a place alloted for them, while the subject assigns his own place
to himself. This distinction must be brought phenomenologically
into sharper relief.
Ready-to-hand things are discovered "there" and "yonder." A
there is determined univocally only in relation to other "theres."
The totality of distinguishable theres is nevertheless related to a
here, from which things are first ordered and space first articulated.
The here has the notable determination of being singular and
non-relativizable, although in a specific sense it is freely selectable.
While the places of the space of action are all distinct from one
another, they are equivalent among themselves: altogether they are
the places of something useful. The of the lived body: is the sale
th_a!jsnot an from
each implement appears as located "there." Here and there are
essentially distinct; there and there are interchangeable, here and
58 Lived Space
there are not. In the space of action, the here is an incomparable
locus; it is the center-fromwhichitis-what itis:-Hielocus ofthe
act!ng subject wh.cij'r.QJi}lisj)Iace, of action-.--
The non-equivalence of here ancr there is --the basis for the
non-homogeneity of the space of action. This is its constitutive
property which cannot be eliminated without eliminating the space
of action and the acting subject himself. In this activity the subject is
certainly not aware of the role of his locus. His primary orientation
is not to himself and his lived body but to the things; the primary
\vQ reference to his body fro_rg!_hings. In the attitudeofacHvity;
the there is prior to the here. If the latter comes thetically to
awareness, it does so inevitably from there and from those indices
that dissolve typical modes of comportment, disrupt the original
07
activity, I can become positionally aware of
my locus when the front of my car has come too clase to the edge of
a precipice; it can become an occasion for my anger when my train
departs from another platform while I am still waiting for it "here."
Although the here is the point of reference in the space of action,
it is exclusively self-related, an absolute here to any there, incapable
of becoming a there in all there-relationships. It is the locus never
abandoned by the subject, who always incorporates it. While moving
toward there he never makes his own locus into a "there," but
J always takes up a new here. As a physical body, the subject is "in"
space just as things are; his locus is thus also a there-yet he is not
1
grasped as a physical body but as the of the space
1 -y1\)t:t of action. Although in the first respect it is a thing among things, in
the second aspect it is irrevocably counter to all things. Thus there
arises a continuous ambivalence of the subject's situation-the
ambivalence of being a physical body and yet being beyond the
physical body. The ambivalence is reflected in the simultaneous
non-relativizability and free selectability of the here.
The acting subject is his cor12oreity; he cannot choose to "have" it
or is not justa corporeity; he is not ananimal completely
caught in, and never capable of, transcending his corporeity.
23
The
23. French existential philosophy here employs etre as a transitive verb
(Marcel, Sartre). Thus the opposition to the older conception of the relation-
ship between physical body and lived body is placed into sharp relief. We
remain with the usual use of the auxiliary verb, the only one customary in
German, chiefly because its transitive meaning is fully understandable only
within contemporary French philosophy, which we are not following here.
Concerning the concept of transcendence meant here, see H. Plessner (2).
The Space Of Action 59
human body is essentially only a surpassed body; the subject is a
body only to the extent that he has it-and has it at his disposal in
the framework of his projects. This simultaneous being and havil}g a
body aQpears in its most originary form in bodily movement. The
subject, caught in the here, nevertheless surpasses his here. While
moving himseff in his nmches the things there, in the world.
The phenomenon of movement will require special treatment. At
the present we are concerned with grasping the center of the space of
action in its specificity-as the place the subject chooses "for" his
body. The entire problem of the localization of the subject in.space \
is found here. While the body in its c_orporeal limitation may_find
itself "here," where am 1 who is constantly "beyond" my b2Qy? The
turn of phrase can only be metaphorical, an image for my being able
to orient myself toward the world, and this orientation is not
but intentional. The being beyond the body ofthe ego does not mean
r a outside of the lived body. Thus we cannot ask
"where" one is to seek the subject who transcends the body; the
question must rather be phrased: how does the lived body of a
subject capable of
Ontologically speaking, the being-spatial of the corporeal subject
has a dual aspect: on the one hand he is oriented toward a there, and
on the other, he finds himself at a there, "being exposed." That -t\.i

which is ready-to-hand is not only a useful means for an end, but is c-JtJ
also highly resistant, detrimental, threatening. The lived body is not
1
merely a condition for the possibility of seizing the world-in its
vulnerability it is constantly exposed to the danger of being seized by
the world. Although a thing aman,?. the corporeal thing is
nevertheless located on the 'tJlde of the thing-world. The
subject is not only a body, but has it and must have it. This dual
determination of the lived body is the reason why its locus in the
space of action is graspable only as a region.
While the subject determines his location by intending the secu-
rity and maintenance of his own lived-body through implements
ready-to-hand, what is more important is the manner and mode in
which this is accomplished. Reduced to vital activity, he may appear
with reactions of flight, defense and protection, such as are also
known in the kingdom of animals. But what characterizes the human
subject is his knowledge of how, in principie, to transform these
Plessner sees it in two ways: "beyond" the body and "into" the body.
"Positionality" is offered as the decisive category of an organic body, while
"eccentric positionality" belongs to a thinking being.
60
Lived Space
reactions through instruments. An implement serves not only to
subjugate the resisting world, but also to surpass the subject's
corporeity, its limits and fragility. In the project of activity both are
"taken into account" and deliberately "calculated."
The use of an implement is informative for the meaning of the here
and the there. At first, the implement is suitable for the progressive
structuration of the regions, the more exact determination of the
there. Furthermore, the possibility of instrumental refinement of
spatial articulation offers a new possibility for the extension of the
here-region. 1t is an extension of a specific kind. lt does not expand
the borders of the here, but incorporates the there into the here.
Something ready-to-hand over "there," out of reach, becomes "here"
as soon as an implement, functioning as an extension of bodily
members, touches it. A dangerous thing situated "there," and
prudently held there at a distance as dangerous, becomes incorpo-
rated into the here-sphere as soon as the armed body can deal with
it. The use of the implement is constitutive for the near-remote
articulation of the space of action; it vares the here-there opposition
and controls the leeway of the acting body.
The interconnection of the phenomenon of leeway with the in-
strumentally refined articulation of the space of action is most sig-
nificant. The reciprocity predominating here can be traced further in
two directions. If the leeway is too limited, if the subject selects his
location "too near" to the things, they will become unhandy or "over-
looked," and their there will remain hidden from his view. Factually
this reveals a structural deficiency of the space of action which can
be eliminated only through the change of place and the here-region.
This destructuration of the space of action becomes preeminent w hen
things, in their resistance and with their threats, are too near. The
place of the subject can thus become a limitation, dissolving each
there and allowing no "room" for action. The limit case of such
destructuration of space, its elimination, corresponds to the subject's
corporeal dissolution. In the space of action there is not only the
too-dense for the there, but also the too-near for the here. The "stable"
space requires for both, a leeway, a region of free motility; the region
in turn must be regarded as its genuine spatial element.
What is the other tendency in the process of progressive
structuration of space? There is a possibility of an attendant expan-
sion of the here-sphere, an extension of the boundaries of its leeway
') through the incorporation of the there. lt appeared that the function
~ of implements can.Ee conceived as a sccessive differentiation of the
there and the yofer; with the of technology,
\\' ~ ?
q \ a\\{ o. \ ~ " ' ~ -
<V
The Space Of Action 61
the leeway approaches the ideal limit case of mathematics. On the
side of the subject, this would correspond to the ideal case of an
extension of its locus to encompass totality-but this means its
dissolution. A mathematically, completely structured space would
not have a center-as a matter of fact, it lacks such a center and is a
fully homogeneous space. Since in its fine structure it is an ideal,
never attainable, limit case, it cannot be a space of action. The
"point" reached in the space of action is always only a region,
which, while arbitrarily limitable, can constantly be thought as still
smaller. Our suggestions anticipate subsequent discussion: if we can ?
meaningfully say that the being of mathematical space is relative to
the subject, then the latter cannot be the subject in his corporeity, but
'- rather must be "outside" this space.
,f It is decisive for the space of action that it be constituted through
;l a corporeal being. Were it a manifold of points, with its center
r it would no longer be toa corporeal being. As
, bemgs, we know no other spatlahty. Nevertheless, we recog- J
nize that in his activity the subject employs something other besides
9 his own corporeal functions in his modes of comportment toward
f _ the world. Thus his space. Q(g_ctiQJl ca_n_o_nl.y_]:fl._ conceived f!'Q!!!Jhe
), possibility of another the mathematical i;q,_ce._This.is_valid
t]J.e acting ]Jody oLitor_not.
Thus for the conception of his space it is essential to see it from the
possibility of mathematics and the exact sciences.
Speaking purely phenomenally, such a mediated mode of being
appears with the things of the space of action. Each relativization of
the h:tre and the there through instruments may reach a degree that
can simply be equated with a dissolution of the limits of the lived
space. Things beyond the horizon can be influenced and used from
"here" through instrumental means. The significance of the horizon
for a corporeal being will be used subsequently as an approach to
interpreting this state of affairs. This is supported by yet another
phenomenon: with the dissolution of limits there appears the
homogenization of the space of action, the dissolution of the natural
center into the "central," the leveling of oppositional orientations by
the reduction of individual actions to their own corporeal center in
'' automation.''
This is not the place to trace out these states of affairs. We are
merely suggesting that th.{}_ gf actior1 revllaJ; th.. :RJ:{}_s_e.rr.cJLQL
another space that _})_() subject of t)J.e space
oCacHon -afready-has it at his _ back, and is
ca-:.defermi:lled by it froill tf18-cU:tseC
62 Lived Space
Most importantly, we must maintain that the space of action is a
topological manifold whose texture is determined by regions. Spa-
tially speaking, even its center is a regan determined on the one
hand by the limitations of corporeity and on the other by the
capacity of corporeality for transcendence. Regarded ontologically,
it is a space constituted in the project of activity, a space whose being
_ is relative to the situation of the acting subject. The situation turned
' out to be in a continuous crisis, holding open divergent possibilities:
from bodily dissolution, and the complete amorphousness of space,
to continuously refined structuration and stabilization, stemming
from the accomplishment of thinking that has transgressed the body
while remaining immersed in the body's own instrumentality.
4. Movement and Orientation. The Space of Action as
Oriented Space
The non-homogeneity of the space of action is not identical with
its orientation. Usually these two characteristics are not distin-
guished, either because it is assumed that both determinations are
given together-which obviously does not imply their identity-or
because their fundamental difference is not even noticed.
What makes the space of action into an oriented space is the
unequal value of the orientations, i.e., its anisotropy. Because a
centered space is necessarily non-homogeneous, the identification
of both aspects is obviously possible. Closer scrutiny shows that
their relationship is determined by anisotropy. Since the latter is
essentially a quality only of non-homogeneous space, it determines
the differing values among the spatial loci. The reversal of this
relationship is not essential and mandatory; it is valid in the space of
action only defacto and is in no wise perceivable as necessary. Thus
although the space of action of a spherical being would be non-
homogeneous, dueto its center, it would nevertheless be an isotropic
and therefore non-oriented space. The orientedness of space is
determined not so much by its center as such but rather by the mode
and manner in which the subject at its center is capable of articu-
lating space in terms of qualitatively differing orientations in accor-
dance with criteria based on his corporeal organization.
In attuned space, the lived body was given as unarticulated. Thus
we characterized the balancing and attunement of torso and limbs
and the preeminent role of the whole body which dominates
expressive movement and subjugates the projective and locomotive
motor functions.
e pa- "1( '"P (_Jr'.J\1<', (( 1
(D'fVO'\'J \'VII(,.d )_
The Space Of Action 63
In the space of action, the lived body finds itself in a changed
situation. Indeed, seen ontically it is the same, yet it represents a
different sense-content. Though it is still a body with the same
modes of functioning, with the same movements, they are under-
stood from a different world context. In attuned space they were
taken solely in their expressive content; here the sense of the moving
body lies in the aim of its activity, in the where-to of the movement.
Upright movement is typically human. Thus it is mandatory to
relate it to the verticality of space. It is usually taken to be the
direction of the body's axis of symmetry.
Undoubtedly, verticality is a dimension of the human space of
{') action. The animal does not live with things that are located
_ f>' "above," and standing is typically a human bodily posture. Only in
f the human space of action is the vertical a dimension: it is a
continuum of possible opposition of orientation between "above"
and "below."
24
One distinction must be observed. For the sake of exactness, we
must distinguish the designation of locations such as "above" and
"below," "in front" and "behind," "left" and "right," from the
dimensional character of these opposites. Indeed, each designation
of location 5ncludes a relationship of orientation. For the acting
subject, the separation is not to be understood as a distinction in the
modality of givenness, as if at first there were the presence of an
opposition of locations and subsequently there emerged a moment of
orientation. Phenomenologically speaking, this could not be main-
tained. What the analysis intends with the distinction suggested is
rather to understand, in the first place, the extent to which above and
below are possible as opposites and, secondly, how oppositional
orientation emerges on the basis of the opposites.
Concerning the opposition of location between above and below,
the first question is already answered: upright posture must be seen
as the final, irreducible condition for the possibility of this opposi-
tion. But how are above and below conceivable as orientational
opposites? How is it that the acting subject distinguishes between
above and below?
The fact of upright posture is obviously insufficient asan answer.
Moreover, the conception of it as "upright" already assumes the
24. By dimension we understand a continuum of possible transitions of
orientationaJ oppositions; thus not every opposition is dimensional. Fur-
ther, it ought to be pointed out that this concept of dimension has nothing
to do with the mathematical concept ("degree of manifoldness").
64 Lived Space
opposition between above and below. Indeed, the subject relates
both aspects to the lived body-otherwise above and below "in
themselves" would be senseless. This does not clash with the fact
that among all the oppositional pairs, above and below is the most
stable; it is not "taken along with" the body. (This stability results
from the univoca! orientation of falling things, pulled by their
weight.) This oppositional orientation remains the same even in
different bodily positions, such as the horizontal posture. Indeed, in
this case the head remains "above," yet the physical location is
different from the indicated direction of orientation. In addition, this
fact shows that directions are not present only with corporeity and
its organization. If the unequivocallocalization of bodily zones were
possible through sensations, through a "felt sense" of the body, or
with the aid of the body schema, no orientation in two distinct
directions would be given. This suggests that the bodily sensations
themselves must be determined by orientations.
Orientations are neither corporeal nor in or of things. They are
relationships of the lived body toward the thing; these relationships
are neither causal nor telic, but primarily functional relationships
first constituting themselves in the interplay between, the acting
body and the world to be acted upon. And they are given to the
subject in no other way than in the subject's dealings with the world.
While dealing with things I first experience their weight, experience
the "above" as a direction in which I exert my force against their
weight and the "below" as a direction in which 1 follow their weight
or the direction "wherein" I must bend in arder to lift something.
Obviously such oriented dynamisms are variations of standing,
specific movements from the upright posture. This is most signifi-
cant for the above-below dimensionalizing of the space of activity;
this interrelationship must not be taken as though something like
orientation is already given by the mere fact of upright posture.
The functional founding of the dimensions appears most strik-
ingly on a pair of opposites whose oppositional character is less
apparent than that between above and below.
The externa! aspect does not offer any criterion for the qualitative
differentiation between left and right. It is not derivable from the
arrangement of the body. This seems to be contradicted by the
longitudinal symmetry of the body, which appears to be divided into
two halves extending equally into two sides. Yet this division is not
something given in the active body; even externally it does not have
the character of a qualitative opposition. This symmetry means
rather a completely externa! equality of both bodily halves, but says
The Space Of Action
P't ,t.P.
r.P \ r \/
fV' V ' ~ , ~ { ~
'o y ./rP )d/ .. \re
\f)" \f { . L' ~ 1 ('
rrr.t') rlp 1 (' 1 ' o:r1f' e(;''
t (!{{' ') e if ,; j-.'> 6 5 '&} ur
nothing of right-left dimensionalizing. From mere image of bodily
appearance it is quite thinkable that in the activity of grasping,
contacting, or touching, both hands may be used at the same time
and act in a mirroring fashion. Such a two-handed being could not
construct a left-right dimensioned space. The left-right differentia-
tion does not inhere in the visible symmetrical physical features, but
in a functional asymmetry of the lived body:t'To "look at" my hands
asn1emnersofmyEody is to find two completely, equally formed
structures; for the eye alone, the one is distinguished from the other
in a freely reversible relationship. It is quite another matter when j
they are no longer experienced as parts of a physical body, but as
functioning organs of the lived body! Seen, it is merely either the
"one" or the "other" hand that grasps; in activity, the right proves
the more active, the stronger, while the left is more unskilled and
weaker. The functional non-equivalence of both hands is the origin
of the qualitative opJ>Dsition between left and right.
25
(It is quite
remarkable that the lived body is asymmetrical only in its hands, the
specific organs of its activity. All the sense organs are not only
symmetrical in structure and arrangement, but also in function. This
will be taken up in greater detail in terms of the space of intuition.)
The individual differentiation in the preference of the one or the
other hand seems to demonstrate that the differentiation can offer no
support for the right-left dimensionalizing of space, although the
latter remains true for all human corporeal beings. Yet this objection
does not say anything against the functional non-equivalence in
general; it reveals precisely what is at issue here. It is not because the
right hand prevails as stronger "for the most part" that it is
significant for our investigation; this can only be attested to by
experienc' What is decisive is that the right and the left hands
25. There is a striking pathological discovery: in the impaired ability to
achieve left-right differentiation symptomatic of apraxia two-handed activ-
ity is not disrupted. (See P. Schilder (1), pp. 44ff.) In the anthropomorphic
region of activity the latter represents a reduced activity conditioned by the
loss of a function. That in all cases there appears a functional disturbance of
the body with a felt disturbance of practical spatial orientation in the
left-right dimension is comprehensible from the interrelationship we have
delimited between bodily function and the space of action.
26. That the functionally higher value of the right hand is not all-
pervasive constitutes a problem for individual research. We may forego the
explication of the developmental researches (Pye- Smith, E. Stier, A. Weber,
etc.) as well as the attempted explanations of physiology, which reduces the
functional asymmetry of the body to the asymmetry of the inner organism.
66 Lived Space
generally act as independent functional members of the entire body.
In human activity the left "need not know" what the "right is
doing"; the lived body is not only both-handed but two-handed.
In any case, the states of affairs here are more complicated than in
the above-below dimension, which is determined by univocal and
non-interchangeable movements proper to it. The opposition of left
and right is subjective in quite another sense than that of above and
below. The latter remains the same in all bodily positions. Positional
alterations of the body are characterized precisely as deviations from
this directional opposition. It is not so with right and left; there is no
bodily position that would deviate from the right-left dimension.
This opposition remains constant. Right and left is meant "from" me
or from someone else. Befare we discuss the orientational moment
inherent in the "from," we must first become clear about the specific
lived bodily situation of the one who enunciates this orientation. It
is only with regard to him that we can obtain orientational univocity
of right and left.
In the space of action the lived body knows how to seize things in
their utility by complying with a principie that dictates its move-
ments. The physiology of movement regards it as a principie of
minimal tension or the minimal expenditure of energy. But regarded
purely phenomenally, this turns out to be nothing else than an
obvious functional preference expressing itself in the maintenance
and preservation of a favorable bodily position. This preference
illuminates the passive role played by the torso in the space of
action. In the movements of practica! activity, the torso appears to be
only a part of the body, which follows the seemingly independent
actions of the organs extending from it.
It is in light of the principie of economy that right arid left first
become univocal specifications. That something is "on the right," I
confirm by certain movements characterized by that principie. Thus
"right" is the appropriate tor pointing or grasping. Al-
though my left hand is not hindJreH'; only my right hand can grasp
successfully and without effort. Yet if specific motives call for the
activity of the left hand, thus disrupting the principie of economy,
the body will pivot immediately and "involuntarily," moving what
is found on the right into a new front-back dimension.
This is nothing out of the ordinary. Yet more careful observation
The reversed understanding of the inner-body organization with respect to
the unequal functional activities of the right and left hands is offered by F.S.
Rothschild.
The Space Of Action 67
shows it to be a most unique achievement. It is not so much that each
right and left are brought into a "third dimension," but rather that
the latter remains omitted. Besides being related to things in front,
the practically engaged human body is also related to things left and
right; it also reckons with those lying "sideward," and knows how to
get them in its grasp without turning the torso.
While left-right dimensionalizing is grounded in corporeity, none-
theless it is nothing corporeal. The directional moment laid out from
me to ... characterizes the spatially polar relationship between the
lived body and the intended things. By means of left and right 1 relate
myself to them and them to me. They obtain their being on the right
and the left through my designations, yet my capacity for such
designations depends on my being oriented to them. Orientation is
always orientation from ... toward ... If either pole is lacking,
orientation collapses.
The functional non-equivalence of orientation in the opposition
between front and back is more pronounced than between the
dimensions of right and left. The "third" dimension is not only
qualitatively distinct from the other two, but is also the most
heterogeneous.
What establishes this opposition is primarily: the organismic ?
constitution of the body. Although muscles, tendons, .q

joints are so organized as to possess a strong tendency toward frontal
movement, the most decisive factor is the frontal position of the
eyes. While the space of action is distinguished from the pure space
1
of intuition, for which another mode of seeing is constitutive,
nevertheless things in their utility remain equally seen and require
visibility in arder to be operative in a project. Only the frontal sphere
of the space of action is surveyable in accordance with place,
situation, and region, and is open to planning; the back-space, in
contrast, is not surveyable and thus is uncertain, dangerous, and
tricky.
With respect to locomotion, the organismic organization of the
body, with the orientational opposition of front and back, contains
yet another strongly formed functional non-equivalence. The for-
ward movement is unequally favored; the space of action is essen-
tially frontal space. That it is not limited completely to the frontal
plane is apparent from the fact that, to a limited extent, actions are
also possible in the back-sphere. This is present for the length of
one's reach and one's stride and is "utilized" in yielding and retreat.
In any case, this form of movement is realized in stretches that are
infinitesimal in contrast to the anticipated forward stretch. The
68 Lived Space
back-space is not actually given consciously; rather, it is there as a
possibility, andas such it exists at all only in this mode of "having."
If a situati9n demands its actualization, for example, in escape, an
immediate turning around occurs and the space of action is trans-
formed into frontal space.
Though the principie of economy present in the left-right organi-
zation may be disrupted by the turning of the body, the principie is
nevertheless maintained in the dimension that comes into view with
the turning of the body; a longer backward movement would be
counter to this principie. This comparison shows most clearly the
preeminent heterogeneity in the directional opposition of front and
)
back. lt results from a superimposition of two mutally supportive
aspects: on the one hand, the organismic structure of the physical
body, and on the other, the principie of economy of the acting body.
The acting body does not live with things behind it as it continu-
ously does with those on its sides. To be sure, the back-sphere is still
co-given and the subject is not set at the periphery of space as in the
space of intuition. Yet he is no longer located "among" the things as
he was in attuned space; the latter literally pervades even its
unperceived back-field with full effectivity.
The prevalent organization toward the front has a specific mean-
ing. lt constitutes the genuine orientation of forward movement,
continuously discovering a new there and at the same time opening
a new front. The acting subject is a being striving toward a goal. lts
activity consists of planning, taking into account, completing, leav-
ing behind, and stepping forward anew-without hesitation and
. l n without a backward glance.
discarded in the space of action. lt has been subsumed by an
1
_L _L eSSeiiffiiYanf-Drrerted-b1rigana thus it too beco mes an oriented
S l)'' space. In this sense, the specific dimension in the space of action is
. ';1 its third dimension.
lt reveals yet another characteristic. Above and below, left and
r -f.L- .,, right are purely spatial orientations.lt is different with the front-back
::... d dimension. lt derives its significance from the "already" and the
yet," and thus indelibly bears a temporal moment. The space-
time nexus differs here in kind from that of attuned space. Indeed,
like the attuned corporeal being, the acting subject "has no time," yet
with the latter the meaning is completely altered. This "having no
time" is affirmed here in a judgment that reveals that one does have
time at one's disposal. His lack of time does not mean an absence of
thetic time-consciousness, as in attuned space, but is grounded in an
explicit positing of determined time. The objective "time-table" time
The Space Of Action 69
dominates space and regulates the movements of action. In distinc-
tion from expressive movements, these proceed in a chronometric
time, and in their execution are submitted toa measurable temporal
duration. In practica! movements, the active subjeCt appears to be
concerned primarily with time. The road to the work-place is not
five hundred-meters long but eight minutes "away." Thus in the
space of action, spatial and temporal determinants pervade one
another in a unique way: the subject here brings time into space,
although he can establish this connection only on the basis of a
previously achieved dissociation.
The elementary directional oppositions turned out to be condi-
tioned functionally; they are oppositional movements of a corporeal
being who, within his established yet motile framework, was capable
of exhausting all the possibilities of his primitive orientation.
The space of action grasped until now as a manifold of regions
becomes accessible to yet a closer scrutiny. From here each there
will be obtainable topographically by means of these directional
oppositions. Up and to the right, down and to the left, above,
below-each are indications of places for the regions; they are
differentiated less through "mere" topological determinations than
through the moment of activity common to them. Even while meant
as pure indications of location and as an answer to a search for a
locus, they are comprehensible only to a being that can move itself
in their mutually posited directions. Indeed, the actual process of
movement does not inhere in their conception, yet they contain an
anticipation of the co-posited path based on the fundamental capac-
ity of a corporeal being to move itself. The problem of the path or
way will be discussed in subsequent paragraphs. The primitive
topological ordering of regions in the space of action was founded, as
we saw, on the originary functional givenness of the bodily orienta-
tions. In this ordering space appears not simply as a pure manifold
of places, but as an oriented and directionally determined space.
This is tied in with specific problem. Since these elementary pairs of
oppositions are related to and accompanied by the body, are they not
merely subjective principies of orientation, mere means to find one's
way in a space existing prior to corporeity? Does not orienting
oneself mean getting one's bearings in a world already spatially
ordered in its own right?
"The" space is represented in natural consciousness as the
"wherein" of all things; its three dimensions are conceived as being
perpendicular to each other. It seems as if these three spatial
extensions "recapitulate" the movement orientations of the body, as
70 Lived Space
if they corresponded to the anatomical arder of the body. The three
semicircular canals of the inner ear, which are responsible for the
balance of the body, also turn out to be orthogonal one to the other.
Thus we could speak of a "natural coordinate system" borne by
everyone.
In arder to incorporate properly this uncontestable state of affairs
into our investigations, it is necessary to indicate the methodological
position of the particular science from which such statements stem:
anatomy. For it, knowledge of the lived body is impossible. When-
ever it becomes an object of research, the body ceases to be lived
body; anatomy obtains its results from the "dead." While results
obtained by these means can indeed reveal partial conditions of
corporeal functioning, anatomy is in no position to comprehend the
functions of corporeity in terms of a living being's body. Rather, it
approaches the object of research with the knowledge of such
functions airead y assumed. Thus the semicircular canals of the inner
ear are related perpendicularly to one another, but that they condi-
tion the sensible orientation of the moving body in space not only
means that the concept of perpendicularity is already presupposed
and understood from elsewhere befare anatomy can employ it; it also
assumes an insight into the interconnections between the mathemat-
ical structure of the organs of balance and the modality of movement
of the lived body. This is possible only on the basis of having moved
and having oriented one's own lived body. Phenomenological anal-
ysis must not adhere to scientific propositions about corporeity, but
strictly and exclusively to the lived body itself, as it is present in the
immediate comportment toward the world and presupposed in each
special science dealing with corporeity.
A phenomenological viewpoint offers no evidence for the assump-
tion of space prior to corporeity, a space in terms of which the lived
body would have to orient itself. Corporeity would have to perceive
space as a pure dimensional system to which it would subsequently
adapt its oriented movements of activity. If this dimensional system
were pre-given, then an account would have to be offered for the
possibility in principie of the intrinsic orientation being continually
mistaken. That such misorientation does not take place factually
would have to be explained by an accident or by a kind of
preestablished harmony-if indeed there is anything at all to
explain here. Instead, we must seize upon what is purely and simply
pre-given, a spatial arder whose constitutive determinations are
inherent in the lived body itself-which means, however, inherent
in what it is in and for itself, lived body of a world, a lived body
The Space Of Action 71
whose possibilities of orientation can only be understood from its
being with another, and whose very spatial principies of arder are
already principies functioning in corporeityP
Further reflections suggest themselves. The above arguments
appsar to imply that the three dimensions of "the" space ought to be
reduced to the three elementary pairs of opposites. But then will not
the arder of space be left in the hands of the contingence of our
corporeal constitution? More precisely, would not the "essential
insight" into the formal structure of space, which pretends to be a
priori and necessary, be placed on the flimsy foundation of bodily
facticity? The mere raising of the question may give rise to the
suspicion of a reflexive circle: one speaks of three elementary pairs
of opposites and attributes to them a preeminence whose justifica-
Han obviously seems to lie nowhere else than in our unsubstantiated
"a-priori" intuition of space. These and similar questions are fun-
damental. In the investigation of the space of intuition they will
reappear in a modified form, although here they cannot be dealt with
fully. They require critical treatment in a subsequent context.
5. The Problem of the Way
The three elementary pairs of opposites constitute a general
framework, and each specific direction of proceeding can be deter-
mined in relationship to them. Each where and there is discoverable
and locatable through ways "in the direction of" right, left, above,
below, in front, behind ... , although the frontal dimension turns out
to be the genuine extension of the space of action. lt attains its
privileged significance through its forward movement of the body.
Seen in terms of a body at rest, it is only one of the directions
mentioned above, even though it is a preeminent one. lt becomes the
sale direction of forward movement by incorporating directions
lying laterally and in back, for these can be transformed into frontal
directions. The space opened in forward movement is a purely
frontal space. The remaining directional determinations in no wise
27. In the pathological realm, the problem with disrupted orientations is
not in the failure of cognitive capacity with respect to the three dimensions
of "the" space, but in the functional failure to project a field of action. The
deviation as such will not be gauged primarily in terms of objective space,
but rather as a disruption of harmony in the sensible processes of activity.
A.A. Grnbaum's attempt to grasp apraxia as an agnosia of spatial relation-
ships has not been clinically confirmed. See the works of P. Schilder.
72 Lived Space
lose their significance for the topography of its regions; yet the inner
oppositions of the individual dimensions, as well as their non-
equivalence, balance out in frontal movement. This relativizes the
directional oppositions and isotropizes space.
A brief discussion of the problem of the "way" or "path" illumi-
nates the results presented concerning the space of action from
another vantage point and clarifies new determinations.zs The
problem of the way assumes its unique position in the space of
action, insofar as the latter is a manifold of regions. The complexity
of the latter is determined by the differing structuration of this space
in accordance with project and situation. Thus sorne limitations
must be assumed- otherwise it would hardly be possible to say
anything universal concerning the phenomenon of the way in the
space of action.
Since the space of action is structured in accordance with regions,
there are many ways from here to there. They create an ambiguity in
specifying orientations toward something. The acting subject is
capable of freeing himself from this ambiguity in the space of action
insofar as he distinguishes one of the ways; this is what is meant
when he speaks of "the" way. It is very seldom that it is the shortest
j connection of a visual path. In the space of action, the "direct way"
seldom has the meaning of a path along a straight line. What is meant
is tlm fastest wy_to_[.f@_Q_l!AI} ?Jl!Jo,_The greatest speed does not mean
the maximum physical velocity of one's own forward movement,
but, rather, in accordance with the principie of economy that orders
one's activities, the minimum of time required. The way in the space
of action is it in principie
from a connecting line segment in visual space. It is subordinated to
a specific viewpoint selected, which in turn is co-determined by a
kv-fo given project of activity.
mJw This also structures the "byway." This does not mean a visually
\ C! oA 1_' ,yr---_
28. K. Levin has discussed the problem of the way in "hodological" space
with the intention of basing a "vector psychology" on a mathematically
clear concept of orientation (pp. 251ff.). His interesting individual results
are valuable for our investigation, even if we deviate sharply from Levin on
fundamentals. Levin begins by asking whether "hodological" space is
Euclidean or Riemannian. Yet the subject considered by us in the space of
action is still "on the way" toward attaining the possible sense of such a
question in the first place. For his vectors and differentials are not phenom-
enally present and require first of all the phenomenological elucidation of
the variously founded concepts of mathematics.
The Space Of Action
73
longer way. Where, for example, an orientation to an instrument
located "over there" is a deviation for the sake of acquiring a tool, in
the space of action it is a direct way. (In this space the path does not
have any points from which it branches off. These relations of
connection between paths, unlike those in the space of intuition, are
essentially determined by temporal components. A curved path
having a double point in a regard loses this
characteristic in the space of activity. For it, two points of a path are
never "the same.")
We assume a further topological simplification by observing only
those regions that have paths in the above delimited sense, which
need not be the case. Obviously the structuration of space plays a
role here. Unstructured regions are characterized by a lack of any
possible path; the previously offered definition of the nondifferenti-
atability of places is identical with their impassability. After all, each (l
distinguishable there in the space of action is in principie attainable
through paths and passages. That these are different for each
situation, that the best way is always different, is compatible with
their relativity to a given project.
The relationship of a path to a situation brings about further
deviations from the space of intuition. In the latter, place B is
situated "between" places A and C if B can be reached without
changing direction on the path AC. This relationship is valid in the
space of action only if each partial segment of the "best" way, AC, is
also an excellent way. The relationships ordering the space of action
do not correspond to those of the space of intuition. Additional
deviations result from the determination of the counter direction. If
one defines the direction of the way back from B to A, it can deviate
anew from what would, strictly speaking, be the reverse of the
original direction, since the way back, BA, is not necessarily the best
way. (For example, if areas are passable in only one direction, then
it may depend on the condition of the areas themselves, or it may
depend on the situation of the subject, for whom the condition of the
path properly assumes its meaning. Hence there is no relationship of
equality between the ways AB and BA; the counterdirection, i.e. the
way back to be chosen, need not correspond with the reversal of the
direction of departure. The way back, in the space of action, is not
the reversal of the way, but rather the way "back" to the place of
departure; both ways encompass a regan in accordance with struc-
tural relationships, while in the space of intuition the way there and
the way back merely constitute a line segment between two points.)
lt must be observed that the best way also contains the moment of


t
.;-v,rw-L
N'\<>kiNJ>

(
74 Lived Space
inexactness and ambiguity constitutive for this space. The region is
the smallest topographically graspable element of this space. In the
space of action, all the places within a region are indistinguishable,
yielding for this space a specific concept of the sameness of a place.
The same can be said of the paths and directions to be taken. Both are
fixed only up to the regions of departure and goal; they are
determinable univocally only if all paths are seen as equally running
from sorne place in a region to the region of the goal. If one
designates the multitude of such topologically equivalent paths, a
domain of paths, then it is true that the space of action, as a manifold
of paths, constitutes a topological manifold.
The definition of equivalence of the space of action, which is
based on topological equivalence, is significant for the determina- ;<
tion of distance in relation to the "best" way. Yet at the same time
the space of action ceases to be a mere topological manifold, since
the distances are determined only imprecisely. The size of a pace
and the extent of a reach are derived from cor oreal orders of
magnitude; they vary with the lived body and are determme
situationaily. We are nevertheless confronted with primordial deter-
minations of measure. Distance is a nontopological concept.
29
The fact that the remoteness of A from B can be quite different
from that of B from A is most significant for the problem of measure.
There is nothing unusual in this as long as the fundamental
difference between distance and remoteness is not observed and one
does not exclude, meanwhile, the possibility of the "later" sciences
of measurement. Their achievement is to be sought in their ability to
change the remotenesses that can only be traversed into measurable
distances. This results in a quantity whose validity no longer
remains related toa subject who is here and now. Rather, they make
it into a pure relative position between indifferent and equivalent
29. This does not mean that it must already be seen as belonging to
geometry. It is distinguished from an exact mathematicalline segment by its
constitutive inexactness and its changeability in terms of the situation of
activity. It is only in accordance with the idea of measure that it is present
as a mathematical quantity. Thus such expressions as the amount of land
covered, or "in a day's work," in a "bushel-sowing" of seeds, by a
"morning's" plowing, among others, are used even today as estimations,
although they can be given in a geometrically determined measure. The old
measures are more meaningful and expressive for a people of a nation.
Besides, they show most admirably the moment of action and the temporal
components of measurement. On this issue, the work of E. Fettweis is most
instructive.
The Space Of Action 75
places. The geometrically dis!_ance is a transformable
quantityThat can only be conceived in
space- Wli_"t tli-"clifiisC-
that has the distinction of being the place
1
of the subject. It is utterly untransformable: it is not a universal
relationship of measure between two indifferent "theres," but a
l strictly singular situation between me in my being "here and now"
.. and the things to which I orient myself .
.JJ' In the elementary space of action, there are only such remate-
nesses, which are eliminated and projected anew by one's own
corporeal movements. There is only a directing oneself from a here
-i .. to a there, never from there back to here: it is always from a new here
..f"to a new there. Thus if the space of action were to remain open only
to the measure of paths and passages, then it would contain only
remotenesses but not distances.
To what extent such a spatial form is lived by primitive peoples
-{ cannot be investigated here more closely. For the space of our 0---
culturallevel, it is only a fiction. Even for the great floods of the Nile in
) ancient Egypt, the riddle is manifest of how this space, while being
space .of also be come space of it
1s poss1ble to geometry to th1s space.
30
Ever smce, the
sJ subject activity, sees himself in a position that
{ \\' suggests a paradox. He is a subject of a space that is projected from 1!
unrepeatable situation, and simultaneously he is shaped by
tors that transcend the situation. Here the question reemerges
ncerning the appropriate interpretation of this ''being-in.'' We can
e a position toward this problem only when the problem of space
" 1s fully unfolded.
6. Nearness and Remoteness in the Space of Action
In the space of attunement, nearness and remoteness are charac-
terized by their qualitative difference. This accounts for the fact that
in it there are no distances, no pure measure and size determina-
tions.
The space of action is distinguished from it. In the space of action,
30. The Egyptians and Babylonians did not "apply" geometry but dealt
with measures obtained and limited to the empirical. Such a way of putting
it is comprehensible from the later scientific consciousness; it is only after
the founding of geometry as a free science by the Greeks that the question of
its "applicability" as such acquired a precise sense.
\

,,
76 Lived Space
the near-remote articulation has altered and carries a different
meaning by virtue of the changed situation of the lived body and its
differently constituted possibilities of transcending. The foregoing
discussion has indicated that the space of action is characterized by
contrast to_Jhe
a !!!Ore --
Indeed, in the space of action nearness and remoteness are not
determinations attributable to things in their ready-to-handedness,
but rather relationships to me in which they are near and remate.
From these relationships it is clear that precise propositions con-
cerning nearness in the space of action can be established to the
extent that the location of the acting subject can be determined in its
here. The sense and limits of such determinations have already been
explicated.
To speak with Heidegger, the originary sense of nearness lies in
the proximal ready-to-hand, in the being-to-hand of that which is at
our disposal.
31
This shows a specific kind of determination of size.
Not only is it inaccessible to exact measure, but its standard of
comparison is different than in the space of intuition. Something
me
by me, and this _ again is nearer than something _ behil},d
from fhespcific strUCtraffi'Hhe-spacefaction
as"frontal space. The measuraEle quantity does not-determine fue
distance; rather, remoteness. In contrast to
distance, corporeity is given in it in a doUble sense: on the one hand,
it is a corporeal body with specific extensions proper to it, and on the
other, it is a lived body in an altered situation. Thus nearness and
remoteness of things themselves become fleeting.
In activity, the remotenesses among things that are near are not
thematized; they are present circumspectively, without any deliber-
ate glance being oriented toward them. lt is otherwise when direct
reaching fails to grasp something present "there." First of all, such a
situation points directly to one's own corporeity. In an undisrupted
course of activity, corporeity knows nothing of itself; only the
"critical" situation throws it back upon itself. lt is only when
something I need is "too high" or "too far away" that I experience
the height of m y body or the shortness of my arm. The consciousness
of one's own corporeal extensions is not given primarily through
measurement, but rather it is experienced through the dimensions of
things in immediate dealing with them.
31. M. Heidegger, no. 12.
The Space Of Action
77
Moreover, it is significant that the crisis situation can be answered
by two characteristic modes of behavior. One consists of the creation
of an implement capable of appropriating the distant object. Here it
is less important to include the estimation of the size-relationships
prevailing and the choice of the implement than it is to bring the
remate to nearness or to create new remotenesses through the
subject's removing hi'mself. What in perception is given as a turning
from the intended there is the condition for the possibility of a
relativization of nearness and remoteness which is characteristic of
the human space of action.
An animal has forward movement as the sale possibility for
drawing nearer. Its relationship to the things of its space is always
immediate. Seen anthropomorphically, this immediacy impover- .@
ishes it. The remoteness of the animal space is always only a vital
distance, dissolving and emerging anew in the fluctuation of its own
dynamism and being resolved by the momentary needs of its body.
The human subject does not only bring the remate to the here by
reaching far beyond the compass of his corporeal structure; it is
rather decisive that he knows how to bring the remate near in a
negative act of turning away. The
thiou-gh the specffic-u-se oflmplements, is reflected spatially in the
phenomenon of corporeal turning away and renewed turning to- ..--
ward. This obviously creates the distance from things that enables
him to be a subject.
As a subject, he also specially manifests a second behavior in the
situation sketched above. Its brief discussion leads to a new aspect.
The situation will be characterized as "critical" because it does
not motivate univocal behavior. Besides continuing his activity
through mediating instrumental apparatus, the subject in such
retains the option of breaking off, _ giving up. Which one will be
chosen depends on--a-g-iven. sftua.fon.CWnaf s essential is that the
latter possibility can call forth a new attitude: instead of indicating
one's own corporeal limitations, there can be a specific orientatioh
toward the remate thing, not grasped circumspectively, but discov-
ered as a thing in pure sight.
This change of comportment is clearly marked in the reduced
corporeal dynamics. The hand, as a grasping organ, is set out of
function and orients itself only by pointing to the unreachable thre:
"the there" becomes desired, or the arm can become absorbed into
the trunk while the gaze "tenses" inquisitively to discern how the
thing escaping one's grasp "really" looks.
Suddenly the thing emerges as apure object. This "mere gazing"
------ -
utfJ.-)1\)
\
78 Lived Space
is constitutive of a new objectivity. Indeed, in the process of
manipulative dealing witntliliigs;lliey are seen, recognized, and
recalled, but their specific perceptual qualities remain behind the
qualities of their applicability. The activity skims over the pure
whatness of things toward their purposiveness; it surpasses them
toward others and toward their context of application. This is the
calm glance attaching itself to a thing that first discovers the
individual in its proper species. Thus simple, fixing observation is
equal to the cessation of activity, and, conversely, the situation of
activity can motivate this new mode of seeing through its specific
structure.
The corporeal manifestation of this structure is found in the
posture of pointing. Indeed, the lived body is more actively disposed
than in the mere viewing, since it maintains the proper organ for the
appropriation of the world. Yet the transformed function of the hand
clearly represents the change of one's position in face of the there. It
rests, and while resting, it fixes the object without reaching it-it
remains empty and reveals a specific possibility of the subject's
being; in its emptiness, it lends "free space" for other modes of
apprehension.
1
The animal does not have hands, it does not point. lts grasping
1
organs are always filled with the proximate and the needed, which
it must appropriate. What is remate for it is only a vital distance. The
hand can be empty, and it would not be wrong to say that ontologi-
cally it is empty, i.e, it must be understood from those activities and
orientations of the subject that allows it to remain in its emptiness.
Moreover, pointing is comprehensible only for a corporeal being
who is with occurs-oiily-hi Tne presence
the sense of pointing must be understood
from the o_ther ubtec::ts. A solip-
sistic sub}ect does not point. P<Enting

/ yi!Y__gf _
- In pointing, corporeity surpasses what is palpably near by; it is at
a there of its space of action yet extricates itself from it in turning
toward it anew. Insofar as corporeity does not reach for the there
actively but lets it remain without any purpose, it posits the remate
as the remate. Pure sight constitutes a different space.
32
Yet the
32. In clinical observations K. Goldstein encountered a very "remarkable
difference" in cerebellum patients who could grasp correctly but were
mostly wrong in pointing. Corresponding observations are in P. Schilder (1);
see also the work of J. Zutt. It is to be noted that similar failures of
The Space Of Action 79
near-remote articulation of the space of action is not different to
sight. Each being beyond the graspable nearness does not only mean
a distantiatioh as a "posture" on the part of the subject, but also and
above all, asan extension beyond into remoteness no longer reachable
in activity. Indeed, the possibility of the relativization of nearness
and remoteness through implements does not eliminate in principie
the horizon structure of the space of action; all activity leads
somewhere to the limit region beyond which space is only space of
intuition. Phenomenally speaking, it constitutes at the same time the
outer border of the space of action lying beyond the project of
activity.
It should not be overlooked that the relationship between the two
sp.aces.is. det.ermi.ned b.y th. e subje.ct_ .. In n .. .. j
active, a_warf1,_ :wh]B-in rmnqte-region
theoretically positioned being, precisely because his corporeal orga-
nlzfttlon does not aow -him to be imything But as the nalysls
cift:Ii can also become a pure
object in pure vision and is thus extricated from the space of action.
It should have become clear that the choice between these two
possibilities is at the same time a choice between two modes of
comportment, or attitudes in the phenomenological sense; each
constitutes a differently modalized objectivity.
Their distinctness appears yet with another phenomenon, to be
understood as the most controversia! in spatial theory: perspective. y-e M:_
Usually it is explicated only in terms of the space of intuition. While
perspective is taken up here, we do not intend to offer its complete
description. It will become clear that the understanding of perspec-
tive is possible only to the extent that we can achieve an abstractive
separation of both spatialities and can also subsequently make their
unity.
In the space of action "there is" no perspective. It first becomes
phenomenally attainable with the thing of intuition, not a useful
thing. The circumspective vision overlooks light and shadow, per-
spectival foreshortenings and intersections, and the absence of
all-sided visibility; such vision transcends them "in view of" sorne-
accomplishment are paralleled by a lack of cognitive achievements in the
space of intuition. Goldstein stresses that pointing and grasping present
activities that are different not in degree but in principie, corresponding to
completely distinct modes of comportment. He suggests that this distur-
bance is a "lack of an objective space confronting the subject," which is "not
required for grasping" (p. 456). This confirms our exposition.
80 Lived Space
thing else. Furthermore, even the size and form of things in the space
of action present determinations that are not thematized in the sense
in which they are for the latter. In activity they are not conceived as
primary qualities, but are incorl!orated in the
is the -J!.k.!'l
of ... ": its form is "appropriate for ... "; it is surpassed things. Thus
of offering themselves
as equal or similar to one another. These relations are not strictly
morphological; they are bound to situation and application and
include relations other than formal ones.
The phenomenon of size constancy belongs in this context; seen
strictly perceptually, it would be a disruption of perspective and
thus would become problematic. lts legitimacy is not to be estab-
lished in the space of activity merely because in near-space the sizes
of things are submitted only to small variations. This argument
overlooks the fact that the near-space, as a zone of action, possesses
--- a differently constituted objectivity than that of a purely visual
region, and it is projected from another center. While acting, the
subject does not see himself in opposition to the world of objects, to
be determined from a specific position in terms of form and size: he
finds himself with a world of usable and resisting things in a
situation in which the individual is continuously surpassed toward
the goal of the actions. Size constancy does not dominate the space
of action because the distances of the near-region are only imper-
ceptibly differentiated, but, rather because in it size as a visual
datum is not at all thematized.
Frthermore, there is a general question: why is it that size
constancy dominates the visual near-region? lt is well known that
perspectiva! changes of the sizes of visual things are not constant; in
a specific near-region (ca. 500 m.) there appears the remarkable
phenomenon of size constancy which is incomprehensible in terms
of the structure of the visual organs and the nervous system or the
laws of light optics. To explain this, the psychology of perception
has recourse to supplementary contributions of memory and sup-
ports itself only with objects one is acquainted with. To bring our
investigations into conformity with the indicated experimental
discoveries of psychology would require specific analyses. Although
in this context they are not required, we shall offer a few suggestions.
Size constancy appears with the perceptual objects of the near-
region because even if they are thematized as pure objects in pure
vision, they are conceiveQ_in !_h_se_ciimented mode of
the ready-to-hand. Here pure contains a viriua.l -momeil.t
The Space Of Action 81
of circumspection of dealings with things; thus size constancy, as a
pure datum of vision, can only be understood from an originary,
active relationship of the subject to things.
7. Summary
The previous observations concerning nearness and remoteness in
the space of action were limited to the corporeal subject in the
framework of his immediate comportment toward the world. They
took their departure from things located "proximally" in the sur-
roundings and thus from a subject as he appears in his immediate
and active corporeity. Yet a remarkable mode of comportment
toward the world also carne into view: the circumspect immersion in
the world of ready-to-hand things was abandoned and the things lent
themselves to pure sight.
Ontically speaking, this is a late attitude. The question is, how-
ever, what is its ontological status? Speaking ontologically, the
relationship of the two spaces discussed here already excludes the
notion of priority of one over the other because in the center of the
space of action there is a being who already has the pure space of
intuition as a possibility. Seen more precisely, this modality of
comportment toward the world is an orientation that is not only
given in specific possibilities of being, but is also an already
actualized possibility and as such ca-determines the space of action.
In our opinion, a limit of the existential-ontological interpretations
is found here. The limitation of Heidegger, and of the French school
that followed him, to Dasein's "circumspectively" being absorbed in
the world-coupled with the claim that the human is thereby
grasped in his genuineness-allows only the space of action to be
considered as existential space. This led to that strange curtailment
of the problem whereby a differently structured space was no longer
a tapie for philosophical investigation. W e do not con test in the least
that-in Heidegger's terminology-orientation and remoteness are
existentials; yet we take both concepts in a more encompassing
meaning which must be taken in a strictly phenomenological sense.
Regardless of how we may conceive our descriptive analysis, it must
deal with an unavoidable issue. Not only does the acting subject
orient himself in terms of his elementary directional opposites;
rather, in his own "existential" space he also has the capacity to
orient himself even if the possibilities of the elementary directions
are lacking. And he establishes remoteness not only through the
coming-into-nearness of distance but also mediately by the use of
82 Lived Space
technology that presupposes the mathematically oriented sciences.
These do not comprise additional gear, but are based on an entirely
distinct mode of vision no longer subsumable under simple "care."
The latter form of lived space leads immediately toward the crest
of mathematization, without, however, surpassing it. More precise
analysis is needed, on the one hand, to show that it is still lived
space, and on the other, to recognize in it the basis upon which the
edifice of metrics can be erected.
Chapter Three
The Space of Intuition
1. Terminological Clarifications
The space of action as a manifold of regions for the ready-to-hand
does not comprise the only kind of oriented spatiality. The subject
does not exhaust his corporeal achievements in active dealings with
things; his functional space is not completed in its being only a space
of action. Besides being a functional unity of aim-oriented activities, ~
corporeity is at the same time a unity of sense-accomplishments; it is
not only an active but also a sensibly intuiting corporeity.
Nonetheless, this does not prove that in the latter case a new kind
of spatiality corresponds to corporeity. It is conceivable that what is
meant here by the space of intuition is nothing other than the
consciously perceived space of action, which, in its structure, is not
essentially distinguishable from the latter, but is, rather, brought to
clearer illumination through a specific paying attention. The func- ~
tional unity of action and perception seems to favor this view.
Two things are to be noted. With regard to the ready-to-hand, it is
touched, seen, heard, and at the same time it is an object of sensory
perception. In a strict sense this is valid only in a vague sense of the
concept of object. Indeed, the ready-to-hand is never outside of sen-
sory perception, yet for this very reason it never be comes its thematic
object. The perceptual qualities are merely "also perceived" on the
ready-to-hand. This "also-perception" does not intend a consciously
posited presentation of the perceptual object.
33
The emergence of
33. We are using this clurnsy terrn in contradistinction to Husserl's
"co-perception" (Ideen 1, 27, 113). Indeed, the "also-perception" is a
"perceptiva co-appearance lacking particular position in factual existence,"
although it is not identifiable with Husserl's conception. Husserl explicates
co-perception on the basis of the relationship between foreground and
background, where in fact an attentional turning of regard plays a role.
83
84 Lived Space
pure perception from it is not sufficiently accounted for by attention.
Pure sight is not merely an attentional transformation of circumspec-
0 tion; is not simiJlY a ready-to-hand
rather has its correlate in a
This does not say anything about the structural differences between
the two spaces. It is furthermore thinkable that despite the distinct
modes of being, the ready-to-hand and the present-at-hand "fill" the
same space. This means that the traditionally followed method of
representing space through the things within it fails with respect to
Q the space of intuition. The treatment of this space will require meth-
odological reflections. The sense of the being-in of things is here
apparently derived from the fact of the material containment of one
thing in another thing, such that obviously the container conception
immediately emerges. If this were valid, then this type of space would
be completely indifferent as to whether it was filled by the ready-
to-hand or by the present-at-hand.
It can be shown that such is not the case. The present-at-hand also
has its space with its own determinations and structural laws,
through which it deviates in decisive ways from the space of action.
In accordance with our fundamental phenomenological principie,
let us again interrogate the subject comporting himself in space and
attempt to shed light on the structure of his space from the side of a
corporeal being as a sensibly intuiting being.
This requires precursory terminological clarification. The reason
we are speaking of the space of intuition (Anschauungsraum) and
not perceptual space (Wahrnehmungsraum) is not because the latter
term could suggest the conception of space itself being perceived in
the same way as are the things in it. Especially since Kant, this
question no longer requires any discussion. It is necessary to show
(Correspondingly there are further psychologically distinguishable rnodes of
intuition, such as becorning aware with an "absentrninded glance," the
fleeting view in a hasty orientation, etc., analyzed more exactly by C.F.
Graurnann.) The also-perception rneant here has nothing to do with atten-
tion, but is rather a title for a potential perception corresponding to changes
of the total attitude. In contrast, our subsequent discussion of co-perception
ought to be taken in accordance with a second rneaning of Husserl: as
potentially itself present, in the sense of categorial intuition, although
fulfillable only through additional actualizing positions (Ideen, 44). 1t
relates to a horizon of "non-presentive" co-givenness and "vague indeter-
rnination," in relation to the full object within the various series of
adurnbrations. Husserl was not fully aware of this equivocation in the
concept of co-perception.
The Space of Intuition 85
the distinction between sensory intuition and mere perception
purely phenomenologically. A thing of intuition means a "com-
plete" thing in its totality of real qualitative fullness, and particu-
larly in all of its sensibly perceivable qualities, inclusive of the
actually not perceived, "covered," co-given factors. It is the unity of
what is perceived and what is grasped along with it that first grounds
the conception of the "thing" as an identity in the fluctuation of
perceptual manifolds. In this sense the thing is "intuited" as a thin&..
while perception, in contrast, offers only changing views of its sides.
Thus each sensory intuition contains categorical moments; never-
theless, it remains sense intuition, i.e., straight forward presence, "in
person," "in the flesh."
Subsequently, the space of intuition will consistently mean the
space of the sensibly present "in person," although categorically
co-determined things and thing relationships. More precisely, it will1
mean the perspectivally and horizonally limited space related to the
intuiting corporeal subject as its center.
In addition, the concept of perceptual space has its own meaning.
While the conception of space as perceivable in itself, or the naive
theory of a spatial sense besides the other senses, no longer requires
discussion, nevertheless a question remains as to whether among the
sensory functions there are sorne whose objectivity reveals relation-
ships of a spatial kind. If this is the case, then the discussion dealing
with space perception acquires an affirmative, although modified,
meaning. Such a space, correlated to this kind of space-presenting
sensibility, will be designated as perceptual or as sensory-space.
Two such sensory spaces, the visual and the tactile, will be come (, l
significant for our line of inquhy. Tlie not to be t&<c. . (
confused with the space of intuition. Although they share certain
formal elements, it is nevertheless a space of visual things that can be
more closely specified, but not of intuited things.
2. The Space of Intuition as a Phenomenal Multitude of Points
In the space of action, the ready-to-hand has its place. The place to
which it belongs is indicated for it within the framework of a project.
The place of the ready-to-hand can be fixed up to a region, which is
the sole topographically graspable element of the space of action.
In the space of intuition these relationships are different. The
present-at-hand things in it have ceased to have a place of their own.
Removed from their contexts of utility, they stand extricated from
any functional relationships with other things. An isolated entity
86 Lived Space
there and yonder has its location merely as a "position" in space
completely external to it; an accidental, arbitrarily exchangeable
somewhere no longer having any relationship to the thing itself.
In complete contradistinction to a visual thing-which, in all of its
qualities, is a function of a location, changing with its changes-the
thing of intuition remains the same in all changes of location. This
phenomenal disattachment of the thing from its position reveals two
properties of space that first appear clearly in objective space: its
homogeneity and its emptiness. This will be discussed in the next
section of our investigations.
This indifference to location of the object of intuition dissolves the
totality of the spatial texture. It is striking that, in the attuned space,
a thing may lose "its" place and be able to change the encompassing
atmosphere; in the space of action this is so much slackened that it
is only a matter of a change of region, while the location of the
ready-to-hand is freely chosen within it; the space of intuition
remains completely untouched by the movement of things existing
in it and remains indifferent to any change.
Indeed, within the space of intuition, regions can be delimited
consisting of parts of space and united under sorne point of "view."
Yet their relative closedness no longer rests on the factual connec-
tion of things, as was the case for the regions of the ready-to-hand
things. The basis for this lies in the changed context of things. In
place of the functional context in the space of action, there appears,
in the space of intuition, the differently structured causal nexus.
Thus its division in accordance with regions is more or less arbitrary.
The regions here are no longer members of a spatial whole, but
merely parts of a space capable of being thought summarily as
contiguous.
However, the decisive distinction appears in the meaning of the
phenomenon of "play-space" or "leeway," and thus in a different
degree of structuration of the space of intuition.
Play-space was defined as the smallest region within which the
place of an implement is freely variable without ceasing to be the
same. Its determinateness contains possibilities of displacement
whose extent depends on the degree of structuration. It was already
pointed out that as play-space of circumspection, this leeway
remains hidden and it is first revealed by pure sight. But this means
that in truth it is a play-"space" of the space of intuition; it is a part
of it and, as a part, accessible to further intuitive differentiations.
The space of intuition reveals a refined structuration unattainable in
the space of action. The process of nesting of regions can be
The Space of Intuition
87
advanced much further than in the space of action, and the sequence
of regions of the space of intuition converges toward a fixed place as
a limit value.
Yet the nesting here is not reiterable arbitrarily. The intuited object
may become smaller and smaller through sorne influence or other,
and its place in space may shrink increasingly, yet this process is not
protractable "into infinity" in the space of intuition; rather, it
reaches a finite limit no longer surpassable by intuition, a limit
beyond which the reiteration can be arbitrarily protracted in thought
but cannot correspond to any sensory intuition. The supercession of
this limit leads to an abrupt disappearance of the object and its place.
If one designates the given convergence value of a nesting of regions
in the space of intuition as a phenomenal point, then the space of
intuition presents itself as a manifold of phenomenal points. As
positive limits of the nesting of regions, they are the smallest parts of
this space, its ultimate topological elements.
Due to its limited approximatability, the phenomenal point is
always a very "rough" formation. The refined structure of the space
of intuition is dependent on a series of factors conditioned by the
organismic order of the corporeal subject-such as the threshold of
stimulation in physiology, the possible impairment of the sense
organs, etc. The degree of exactness is always an empirically
determinable quantity and is the object of experimental researches.
The space of intuition is completely filled by phenomenal points.
It constitutes a (simply) connected set of points whose elements lie
"thickly everywhere."
The concept of thickness in the space of intuition is .different from
that of the space of action. The possible "too-denseness" of the places
appeared in a lack of leeway when the ready-to-hand ceased to be
handy. The space of intuition has no analogue for this. The intuited
is never too dense; it is where it is, as intuited. Its "play-space," if it
has one at all, is one of movement, although it is not as though the
intuited object required a "play-space" of movement in order to be
intuited, in analogy to the way in which the ready-to-hand requires
its play-space in arder to be handy. The play-space is not a
constituent for the present-at-hand, as is the case for the ready-to-
hand. The ultimate topologically graspable factor given in intuition
is not the region, but rather the place, the phenomenal point;
"behind" it there is nothing that could offer a conception ofitas a
region of another extensive manifold to be occupied in a way that
would be relative to a corporeal subject. Indeed, its further
structuration can be thought up to a mathematical point. But though
88 Lived Space
its intuited elements lie thickly everywhere, the denseness of the
phenomenal manifold is not that of the mathematical continuum.
This process of further nesting not only surpasses the limits of
sensory intuition, but leads mathematical thinking into difficulty,
even into an aporetical situation. Underlying the mathematical
concept of denseness, of the multitude of points, there is the problem
of the continuum. Its mathematical discussion cannot be offered
here ahead of time. The sense and nonsense of a mathematical
"continuum of points" can be discussed subsequently in a pertinent
place. Yet it cannot remain unimportant for mathematics that, and in
what manner, the concept of the continuum has an intuitive signif-
icance prior to any scientific treatment.
In addition, it must be strictly observed what really lies in the
unqualified intuitive conception of the continuum-and more pre-
cisely what does not lie in it. Above all, it must be stressed that for
sensory intuition the continuum is not a problem but a fact. As such
it means that all parts of space are interconnected and without
gaps-that, in accordance with a radical formulation of Husserl,
space can never have a hole.
34
The expression loses its appearance of
naivete when it is understood from the givenness of natural intuition
by which it is motivated. This givenness does not possess points of
space but things in space in their own limitations and freedom of
movement. A directly perceivable movement is for the natural
consciousness-which has not yet reached the level of reflection
leading to the Zenonian paradoxes-the originary paradigm of a
continuous event. As long as a thing moves, it is a moving thing in
ea eh phase of its movement. Yet grasped as a moving thing in space,
such movement, apprehended as a change of place, implies that
space itself must be continuously connected. The change of location
of a thing brings about at the same time a displacement of its limits,
i.e., phenomenally speaking, of its surfaces, edges, and comers.
However, these limit formations constantly and necessarily coincide
with the corresponding spatial formations. In particular, each comer
point of a moving object continuously coincides with a phenomenal
point of space, and, conversely, each phenomenal point of space can
in principie be a comer point of an object.
It is the latter possibility that is primarily responsible for the naive
conception of space as a (phenomenal) "continuum of points." It
means only the intuitive possibility permitting any arbitrary place of
space to correspond to a comer point of a thing, thus signifying in
34. E. Husserl, Ideen II, 13.
The Space of Intuition 89
this sense the division of the spatial continuum into phenomenal
points-more precisely, its successive divisibility into mere individ-
ual theres and yonders. This conception does not imply that the
spatial continuum "is" an (actually given) multitude of points and
"consists" of mere points that when combined constitute space. In
natural space-consciousness what is present is not what the contin-
uum is, but only what can take place within it. Strictly speaking,
there is no "atomistic" continuum as a given in natural space of
intuition. Rather, there is a theory of an atomistic continuum derived
from a conceptually reiterated extension of the intuitively given; the
paradoxes of this theory are grounded on the latter (see p. 296 ff.).
3. The Lived Body as the Center of the Space of Intuition
As with the space of action, the space of intuition has the
corporeal subject for its center. Thus purely formally this space
possesses all the properties resulting from its being centered. The
singularity and uniqueness of the here, and its radical incompara-
bility with any there, constitute the nonhomogeneity of the space of
intuition. They articulate it into nearness and remoteness and
condition its finitude and horizonality.
The situation of the lived body in the space of intuition is different
from that in the space of action in two respects.
In distinction to the space of action, the space of intuition is
present as a space of remoteness. This determination also has a dual
meaning. First, it means a spatial remoteness between the things and
myself. In this sense, its remoteness in fact surpasses that of the
space of action, although the latter is also characterized by an
articulation into nearness and remoteness. The second meaning
consists of a characteristic remoteness of the body in relationship to
itself. Indeed, in the space of action the lived body is not present to
itself, but is rather at the things, is oriented toward them; yet it has
them only in immediate contact with the body. In the space of
intuition the body is beyond itself in an entirely different sense, and
this by virtue of the distinctness of the visual function which plays
a decisive role in sensory intuition. To pick up something ready-to-
hand and to "fix one's eye upon" something present-at-hand,
grasping it visually-here not only two fundamentally distinct
attitudes of the subject, but also two entirely distinct relationships of
the relevant thing to corporeity are manifested. The latter expres-
sions are meant metaphorically, since the thing has no way of being
in the eye as it is in the han d. Vision do es not occur through the body
90 Lived Space
as corporeal in the same way as with handling. Indeed, the eye is a
member of the corporeal body, insofar as it is sensitive to touch and
pressure, but not insofar as it sees. As seeing, not only is it not given
as it is when seen, but it cannot be given-as seeing-through any
kind of objectifying observation. In seeing it determines itself not as
a member of the physical body, but as a corporeal organ in function.
And since this function is an objective one, it means that the eye
does not have its object on or in itself, but in space. While intuiting,
corporeity appears in a mode of functioning that has its fundamental
conditions in its own body, but does not bring corporeity itself into
view. The higher value and the higher power of achievement of the
visual function lie in that the intuiting corporeity looks away from
itself, since it can effectively assume its function only in a specific
spatial remoteness from corporeity.
This transformed role of the lived body in intuition is presented in
bodily corporeity itself. Its own dynamics are reduced qualitatively
and quantitatively; they sink to a minimum to the extent that those
movements in which corporeity appears as active, in a narrower
sense, become redundant for perception. In this space, limbs and
torso are deactivated in their specific functions and remain motion-
less as an undifferentiated whole; phenomenally, they are nothing
more than mere bearers of the senses. What remains essential in
corporeal movement in this space is, strictly speaking, only the
movements of the sense organs and the equivalent forms of dynam-
ics such as the movement of the head and locomotion. What is
noticeable in intuiting corporeity is above all its organic and
functional symmetry. All organs are superposably related to the
bodily axis, and the functional differences of the sides are dissolved.
How does this altered situation of the intuiting corporeity appear
in its space? What does this situation mean for space?
4. The Oriented Space of Intuition
The intuited space is also an oriented space with its center in the
oriented corporeal being who emanates rays of intuitive intention-
ality to things and encounters them above or below others, on the left
or to the right of them. The topography of this space acquires its
sense only under the assumption of a standpoint; it is a description
of a location in the framework of a corporeal system of relationships
whose point of departure is indeed freely choosable, yet within
which any given location is anchored. Thus the arder of things
among themselves is here not a pure relation of position, not a mere
The Space of Intuition 91
constellation, but rather still a situation-an ordering of the there
relative to the here and now of a corporeal subject who apprehends
what is sensibly intuited. Each relationship of arder obtains its
meaning from its associative relationship. It is not merely a factual
arder of things in their pure places but an emplacement of accom-
plishments of the subject at the present in a here and now.
This system of relationships consists of the three elementary
orientational oppositions. To discuss them in detail once again
would be a repetition of the essential traits delimited in the previous
section. That they were already discussed earlier was not due to an
arbitrarily systematic procedure but, rather, was necessarily based
on the state of affairs: the orientational oppositions are primarily
functional oppositions in a narrower sense; they are distinctions
whose understanding is to be sought with corporeity as an active
subject. The consideration of the oriented space of intuition as such
would be unnecessary if its directional oppositions consisted merely
of those of the space of action. Factually this is the case; the sense of
the directional oppositions in the space of intuition is the same as
the space of action. Yet with the former, certain changes appear.
They will become comprehensible from the interrelationship of
orientation and movement. As has become obvious in the previous
chapter, the directional oppositions are comprehensible only from
the dynamics of the corporeal subject. Each where that is determined
in orientation is bound to the whither of the movement and the
possibility of movement. The contrary directional oppositions result
from the qualitative difference and opposition of the singular dy-
namic forms. In the space of intuition these are leveled out to a great
extent due to the fact that the movements of intuition are reduced to
the minimum. The characteristic feature of the space of intuition in
contrast to the space of action lies in that the anisotrophy of the latter
is here equalized.
This is most conspicuous in the weakly formed opposition be-
tween left and right. As a functional difference, it is completely
meaningless in the space of intuition. The differentiated activity of
the left and right hands in grasping disappears completely in the
space of intuition. When the hand is used to point-in its only
function within the space of intuition, a function intermediate
between grasping and seeing-left and right are equivalent. The
remaining dynamics in this dimension, the movements of the gaze
and the corresponding movements of the head, are externally
equivalent; they are also felt "from within," kinaesthetically, as
qualitatively different.
92 Lived Space
This is correspondingly the case with the pair of opposites
above-below. In the space of action the emphasis is placed on
qualitatively distinct movements engaging the entire corporeal body,
while in the space of intuition we notice only the dynamics of
raising or lowering the glance. The movement is slighter the farther
away the things are and the more clearly the space of intuition
presents itself as a space of distance.
The most pronounced and significant change is undergone by the
third pair of opposites: front and back. While in the space of action
it has a most pronounced opposition, in the space of intuition it is
eliminated-not by balanced, qualitatively equivalent movements
making it isotropic, but rather through the fact that the dimensional
tension is radically annihilated.
Indeed, the space of intuition, like the space of action, does not
completely clase itself off with the frontal plane. The plane can
extend with the "wandering of the glance," and the horizon appears
closed at the limit from which the line turns back upon itself. Thus
the actually intuited part contains a conceptual moment of being a
mere segment. Nevertheless, the back-field disappears-and indeed
on the basis of the exclusively frontal organization of the visual
function. Strictly speaking, seeing "toward the back" is functionally
impossible. Even if we were to turn our head around, considered
functionally, we would still be oriented forward. In this sense, the
space of intuition is always frontal space.
While attuned space is experienced as surrounding, effective in its
all-encompassing fullness, and the space of action allows limited
possibilities for action in the back-sphere, in the space of intuition
the back-sphere is finally lost. The lived body moves toward the
periphery and is no longer "among" the things, but has them
exclusively over against itself. It is no longer located in space,
although as corporeity it is itself spatial. Thus in this placement it
represents a dual polarity consisting, on the one hand, of the
inner-spatial there of things and the here of its position, and on the
other, of the non-spatiality of this polarity, since it is tensed between
the spaceless subject and space itself as an object.
This is a most remarkable fact. The subject has attained a mode of
being that truly legitimates his title as subject: he not only intuits
things in space as pure objects, but also confronts the world as an
object and regards space itself objectively. "The" space means
primarily a space that is regarded as one for all corporeal subjects;
the subject knows that his space of intuition is a "mere segment" of
it. This specific mode of conception is not first introduced into the
The Space of Intuition 93
perspectiva! space of intuition through reflection, but is rather
present from the very outset in the everyday awareness of space. No
priority or posteriority can be demonstrated here. Only a subject who
has the space objectively is at the same time corporeally located at
the periphery of a perspectiva! space of intuition as his own and,
conversely, finds that his perspectiva! space of intuition is not given
to him otherwise than as a "part" of "the" space. It is only by
respecting this state of affairs that we can enter into the variously
stated question of how the corporeal subject can derive a conception
of the one space from the oriented space of intuition given only to
him from time to time. Certainly he does not move from the first to
the second or from the second to the first. His monadologically
spatial world is nothing else than the mode and manner in which he
as a corporea1 subject already has the presence of "the" space.
But this space is no longer relative to a corporeal subject in his
monadological isolation. If it has any relativity of being, then it is
relative to all corporeal subjects. "All" here does not mean a
numerical multiplicity but, rather, allness as a unitary form of
intersubjectivity with a specific consciousness-structure. That the
latter is historical, that the objective consciousness of space is
historically changing, is not being overlooked, although it can be
explicated in greater detail only subsequently (pp. 156 ff.). In any
case, strict reservations must be maintained concerning assertions
about the structure of this space. Such reservations do not imply a
postponement of a problem but correspond precisely to the state of
the phenomenon of the natural space-consciousness. At this level of
space intuition, space is not yet given thematically but is present in
the singular space of intuition in the form of the "continuation" of
space "beyond" the horizon.
The following investigation is valid primarily for the monadologi-
cal space. From the foregoing it is obvious that we cannot answer the
question of how objective space develops, unfolds, etc., from it.
Rather, following our method, it is necessary to show how the one
intersubjectively conceived space appears in the perspectiva! space
of intuition.
5. Spatial Depth and Perspectivity
The corporeal being located at the periphery of space has become
a subject; he finds himself in a position of absolute opposition to the
world as the totality of sensibly intuitable objects. But they remain
only what is present of them "in the flesh": they are related to the
94 Lived Space
position of the lived body, whose outlook toward the world is
monadological.
This intuiting subject has the world befare and for himself, while
the problem of spatial depth appears in the having of the world for
oneself.
How is spatial depth present to me in my here and now? It is not
an object of my intution in the sense of immediate self-givenness,
like the things befare me. Only they are there "in person," are found
there and yonder. (We must not overlook that, strictly speaking, they
are present "in person" only from specific "sides," while the other
sides are only co-meant. At this point we still exclude perspective
and accept the things only in their being behind one another, i.e.,
without regard to perspectiva! changes.)
Between the things of nearness and remoteness there is a charac-
teristic relationship of priority inconceivable as mere differences of
distance between things among themselves.
The things nearby are open to my glance and are given in full
actuality. Their dimensions can be completely traversed from here
with a wandering gaze. This is not valid for remate things. The near
and the remo te stand in a relationship of covering and being covered,
of the hiding and the hidden. This relationship is reciproca! insofar
as the one calls forth the other; it is irreversible insofar as it is
unequivocally determined by the single privileged point of my
space, my standpoint. This space is exclusively frontal space.
However, 1 do not have things only befare me, but also befare other
things. The remate object is haunted by an unresolved residuum for
intuition. The open, the obvious appears on the dim background of
intuition of the alien, the questionable, the undiscovered but still to
be discovered. The hiddeness of a distant object is not something
irrevocable, but a momentary fact. It is conditioned by my being here
and now and can be abolished through locomotor movement.
Locomotion in the space of intuition is motivated solely by what
is hidden, undiscovered; its sense derives from the aim of discover-
ing. In its being still undiscovered, the present-at-hand is the proper
object of the self-moving corporeal subject in the space of intuition.
In covering and being covered, space appears in its depth dimen-
sion. It is essentially distinct from lateral extension.
35
Extension has
to do with objects; it is a pure characteristic of things, graspable in
35. The distinction between the surface- and depth-extension of space is
clearly worked out in detail in the little-noted but excellent work of H.
Lassen (1), specially pp. 124ff.
The Space of Intuition 95
pure relationships of size. It is not so with the depth dimension.
Depth cannot be measured as a spatial interval. Measurable dis-
tances are reversible and transposable. Depth, however, is not a
measurable interval but a remoteness, a polarly tensile relationship
between the things and me; it is irreversible and strictly singular.
Corporeally 1 am constantly here and now, a lived body in a
situation. Thus depth can never be abolished; not only can it never
be surpassed, but it is always created anew in locomotor movement.
Therein lies the phenomenal state of affairs: 1 can never wander in
the oriented space of intuition, I always take it with me. In all my
traversals of space it is always befare me. Although from moment to
moment it is different in a continuously changing plenum, it
nevertheless remains structured in accordance with surface and
depth, belonging to me as long as 1 am sensibly and intuitively
oriented to the world.
My stance in opposition to the world and my capacity to turn
toward the world are intelligible in terms of spatial depth. As the
genuine dimension of my being, and in complete ambivalence
concerning m y being as lived body oras consciousness, depth is first
comprehensible in its own equivoca! sense: first, it is the sole
dimension into which it is possible to move forth; second, it is
precisely the dimension that cannot be surpassed in progressive
movement forward. On the one hand, it is a space that requires me
as a motile being to occupy the center, and, on the other, it is at the
same time a space in which 1 cannot wander. This apparent paradox
can be resolved only through an ontological explication of the spatial
depth of the subject- of the subject in his corporeal restriction to the
here and now and in his intentionality of consciousness, which
transcends corporeity. It is important to take the latter in the narrow
sense of an ability to orient oneself in pure opposition to an
object-world, free from the effects of space from concernful absorp-
tion in a merely "proximial" world of ready-to-hand things at our
disposal.
Grave arguments have been raised against the phenomenon of
depth. Berkeley's theory of space became the starting point for a
series of discussions concerning the question as to whether or not
there is an original depth perception, and Berkeley's arguments
elicited a multitude of treatments of this problem in philosophy,
psychology, and physiology.
The central conception of Berkeley is that real space is solely
two-dimensional. It is associations of a determnate kind that first
create the "idol" of the three-dimensional space of depth. In con-
96 Lived Space
nection with Locke's theory of association, the emphasis lies, on the
one hand, on the linking of tactile and visual data; and, on the other
hand, on the unification of acts of perception and thoughts. Thus
through the experiential knowledge of the all-sidedness of the
spatial thing in the individual intuition, extension in depth is
associated with the surfaces. Furthermore, the spatial depth was
explicated physiologically through binocular perception, which
ought to unify stereoscopically the non-corresponding two-dimen-
sional retinal images (a later term). With each of these arguments the
peculiarity of the depth dimension became clear. lt is characteristic
that, during the following periods, one followed one of these three
trends of thought, and that the theories of space evolved during the
nineteenth century were systematically devloped in three direc-
tions.
The last of the arguments presented became the point of departure
for physiological and psychological observations of the senses. Their
results, especially since the discovery of the stereoscope by
Wheatstone (1838), led toa hardly surveyable profusion of individ-
ual analyses and quite extensive experiments, which cannot be
explicated here in any greater detail. What is relevant in this context
is the total methodological milieu of the individual researches
concerned with the phenomenon of depth.
36
In brief, the main
difference between their mode of observation and ours is this: theirs
is concerned with explanation, while ours with understanding. They
inquire into the physiological conditions that must be fulfilled in
arder to result in depth perception: they find these conditions in the
form and arrangement of the visual organs and in neurological
processes. In contrast, we inquire into the sense of seeing for the
subject; we do not take seeing as a process functioning in accordance
with physiologically and physically determined laws of the organ-
ism but as a "function" of the subject, as a mode of his having a
world. The immediate question is whether a causal-explanatory
science, with its constantly repeatable experiments and its verifica-
tion of statements by an individual case, does not deserve preemi-
nence over a point of view having nothing better to offer than sorne
dubious "evidence." Such a question should be a warning that
36. A good overview of the historical development of this research is
given by F. Sander. The work by W. Arnold, the title of which is misleading,
deals with the experimental method of the psychology of depth perception.
For individual problems see N. Guenther, E. R. Jaensch, F. F. Linke, G.
Lintowski, F. Mayer- Hillebrand, and B. Petermann.
The Space of Intuition 97
the fundamental method prevailing here has been misunderstood
and misinterpreted. It considers results that have been disassociated
from the context that grounds them and sets states of affairs against
one another where only methods are to be compared.
The individual scientist must be cognizant that what he seeks to
explain can be interrogated causally only because that something is
in sorne manner already given and open to understanding. This fact
is no less than self-evident. With regard to our problem, it means that
the eye must already see befare it can be investigated by a special
science as a condition of seeing. Seeing as a function of corporeity
makes the very question of the structure and mode of functioning of
the visual organ understandable. The lived body is presupposed by
the organism; it was already a body capable of functioning and acting
befare it could be come a "re" -acting body, an object of research for
the special sciences. What decisively determines their view, how-
ever, is a rigorously controlled methodological procedure that con-
tains measures for the reduction of the corporeal subject to a
"corporeal" object; this is precisely why the lived body as living, i.e.,
as a functioning corporeity, escapes them. The special science, must
necessarily divest corporeity of its situation and place it in a factual
situation that is repeatable at will in arder to obtain its results. Yet it
gains its propositions at the price of the living body. Strictly
understood, corporeity radically escapes any objectifying scheme of
investigation. Irreducible in principie to constant and arbitrarily
reproducible experimental conditions, it is rather already itself a
participant in establishing such conditions. As the lived body of a
subject who visually investigates the visual process in light of
physiology and optics, this lived body stands outside of the postu-
lated methodological scheme and leads unavoidably to a circular
structure of the objectivating observation of vision: observation must
presuppose the capacity that it seeks to investigate.
Obviously the claim of the special investigation is justified if it
does not pretend to be more than an explanatory science, a system-
atic analysis of the organic conditions of vision. To the extent that
the regularities investigated are those of vision itself, the investiga-
tions are legitimate; after all, lived corporeity is not without a
physical body and constitutes an ontic unity with it. Yet such
investigations can never be ontological. They cannot conceive of the
living corporeity as a mode of being of the subject. The regularities
of the latter are different from those of the organism.
Phenomenology and ontology cannot criticize the results of the
individual sciences so long as the latter do not infuse the copula of
98 Lived Space
their expressions with metaphysical meanings. They must conceive
of themselves as sciences of entities limited by a specific method-
ological view. Philosophy remains primarily on this side of individ-
ual investigations. Yet insofar as it documents and also establishes
such this-sidedness, it transcends the individual sciences toward
a philosophy of scientific entities themselves, thus becoming a
theory of science. While it attempts to become an explication of
being, it is mistaken if it pretends in this manner to become a
"fundamental" ontological endeavor. lt cannot simply remain "on
this side" of the individual sciences, unaffected by their accomplish-
ments. If it is to become "fundamental," it will have to take a
position not befare all sciences, but behind them in arder to
illuminate their ontological foundations.
Regarding our problem, we must acknowledge that diagonal
disparity "must be" recognized as a causative factor in depth
perception. That depth perception is possible in this manner does
not, however, affect another way of looking at it: "seen" primally,
depth is a clue for the intentionality of the corporeal subject, its
being as consciousness.
The thesis of the association between tactile and visual data has
frequently been discussed. Among the newer philosophical theories,
Husserl's early phenomenology revives the old Berkeleyan concep-
tion, though in a greatly modified guise, in the theory of constitution.
According to Husserl, the constitution of the thing occurs in various
syntheses, and, above all, in the sensuous syntheses of visual and
tactile primordial objects; such constitution takes place at many
levels.
37
Following the pre-spatial fields of the first level-where the
visual and tactile fields consist of completely separated, quasi-
extensive manifolds in which only the hyletic data are individu-
ated-the second level extends the field through the motorics of
bodily members, in which the visual and tactile data are united in a
determinate manner. Oriented space is constituted only at a third
level. lt is built from the oculomotoric field through "reinterpreta-
tion" of the visual depth into a third dimension. The basis for the
reinterpretation is found in kinaesthesia, which presupposes carpo-
real tactility.
The notion of reinterpretation is based on the view that space is
not originally "given" but, rather, is "constituted," such that unbe-
knownst to naive observance, it contains unnoticed traces of preced-
ing constituting "achievements" of pure consciousness. Behind this
37. E. Husserl, Ideen JI, 9.
The Space of Intuition
99
question there is a fundamental problem that completely separates
the spirits of the new phenomenology. Thus existential philosophy,
specifically the French style, sees its irreconcilable opposition to
Husserl precisely in that for Husserl, phenomenology is irrevocably
bound to the transcendental question. From the very outset, phe-
nomenology for Husserlis constitutive phenomenology, and specif-
ically, it is a "systematic uncovering of the constituting intentional-
ity." The discovery of its typical character of accomplishment
reveals the hidden origins of the objective phenomena in transcen-
dental subjectivity. The object itself is then ultimately only a
"transcendental clue" for the analysis of its constitution in transcen-
dental consciousness.
38
In sharp contrast, the fundamental principie
of phenomenology according to Merleau-Ponty is expressed as
follows: "mon acte n'est pasoriginaire ou constituant, il est sollicite
ou motive." ["my act is not primary or constituting, but called forth
or motivated"].
39
In place of the constitution of objectivities in pure
transcendental consciousness, there emerges a motivation of the acts
by the "in-itself" (en soi), which is neither a constitutive nora casual
process but rather a "medium" between both. This is not the place to
pursue this topic. Here the question is how these two contrasting
conceptions relate to the spatial depth phenomenon.
The phenomenon as such is not the least denied in transcendental
phenomenology. Husserl's methodological reductions reveal that he
intends to interrogate the purely and simply given in order to illu-
minate retrogressively the activities in which the given became es-
tablished as a result. This requirement need not be rejected a limine;
rather, we must ask how such a layered constitution is to be under-
stood, and what kinds of ultimate givens are revealed. For Husserl
these are the hyletic data, which are formed intentionally by the
animating, sense-conferring conception that makes them objective.4o
Yet it is repeatedly objected that these cannot be "phenomenologi-
cally brought to light." However, this assertion is notan adequate
counter to Husserl without further comment; he does not challenge
the immediately given, unreflective attitude. Here, what it means to
bring something to light phenomenologically has been shifted. In
accordance with his conception of transcendental phenomenology,
38. E. Husserl, CM, 41; FTL, 97.
39. M. Merleau-Ponty (1), p. 305.
40. E. Husserl, Ideen I, 41 and 85-88. (In this context it is irrelevant
that for Husserl the hyletic data are not the ultimate givens; rather, they in
turn are constituted in originary time-consciousness.)
100 Lived Space
Husserl is not concerned with the immediately given but with the
demonstration of the constitutive conditions for the possibility of the
given. This demonstration moves at an entirely different level of
phenomenal elements than those his critics presuppose.
If we do not concur with Husserl's conception, it is not beca use we
agree with the arguments of his critics. Rather, the states of affairs
require Husserl to bring to light the content of the individuallevels
of constitution; he must offer us meaningful access to how the
constitution of objects must be thought. It must be possible for the
constitution of objects in pure consciousness to be insightfully
traced by the consciousness which reflects on this constitution. Yet
this demand is plainly not met by Husserl. The hyletic data and the
sense-conferring acts of noeses belong to the really intrinsic [reelen]
composition of the experiential stream, while the noematic object
constituted by them as intentional is foreign to the acts. This
constitutes a sphere of objects whose phenomenal transcendence is
no longer maintained in its originality, but, rather, whose origin is
transposed into the really intrinsic conditions of experience. Never-
theless, such an object cannot be comprehended in this transcen-
dence in terms of its assumed origins. And the aforesaid transcen-
dence itself is incomprehensible when it does not mean a really
[reell] existing object, but only a transcendence immanent in con-
sciousness (incorporated into the really intrinsic immanence of
experience).
41
It appears that no matter how insistently Husserl's distinction
between hyle and morphe is maintained, Husserl himself could not
adhere to it with the strictness and consistency demanded by his
own theoretical strictures. Whenever Husserl purports to operate
with "hyletic" data, it becomes obvious that, strictly speaking, these
are already "animated" contents. Since according to him, they
constitute themselves in originary time-consciousness and are next
to one another and, one after the other, are unities, units, and
multitudes, they are already categorical formations even in a Hus-
serlian sense.
41. Husserl's use of the terms "transcendence" and "immanence" is
confusing insofar as the intentional object is characterized by him with both
terms in relation to the unreduced, to the reduced object, or to the really
intrinsic composition of an act. We shall not trace this problem any further
here, since it hardly concerns our theme. Concerning the critique of the
constitution of objects of intuition in Husserl, see H. U. Assemissen; on this
problem see especially pp. 66ff.
The Space of Intuition 101
As our suggestions indicate, we cannot follow Husserl on these
points. If our reflections are correct, then the pre-spatial field is not
maintainable phenomenologically. The reinterpretation of the visual
field into an oriented visual space, the apperceptional shift from
one-next-to-another and one-upon-another on the surface into
one-after-the-other in space, which must necessarily be introduced
in arder to attain the intuitive spatial depth from the merely
quasi-extensive manifold, seems to be an intellectual construction
whose individual stages of constitution are not clear and which is
inappropriate for the facts of the case. (Indeed, there is an ap-
perceptional change in a reverse sense, namely from space to field;
yet this transformation is only possible on the basis of the originarily
given space. It is a specific, abstractive achievement of the intuiting
subject. See pp. 124 ff.). In contrast to Husserl, we accept spatial
depth as primordial, as a given that can no longer be interrogated in
terms of sorne hidden associative activities.
42
Merleau-Ponty also conceives space in this sense. It is understand-
able that Merleau-Ponty develops his conception of space against
Berkeley, and it is instructive that this occurs in his very first
arguments. For Berkeley the impression of spatial depth results from
the association of the frontal and lateral surfaces of a thing. Depth for
Berkeley is nothing without this association; the depth surface of a
physical body is in truth once again a frontal surface that is seen
from the side. The impression of a thing as three-dimensional thus
only appears when this lateral aspect is co-present in thought. This
argument contains two presuppositions: on the one hand, the
possibility of presenting a three-dimensional thing in a surface and,
on the other, the inclusion of the corporeal capacity of locomotor
movement. Both presuppositions are discussed by Merleau-Ponty.
Through examples of perspectivally presented objects, he demon-
strates his theory of the freedom and facticity of perception, from
which it follows that "la profondeur nait sous mon regard,
parcequ'il cherche a voir quelque chose." ["depth is born beneath
my gaze because the latter tries to see something"].
4
3 His point of
departure is deliberately taken from represented figures. If the
observer already sees a thing as a spatial thing in what is really a
42. An indication of the originality of spatial depth is present, according
to C. Stumpf, in the fact that within an intuitable surface of any kind,
planarity and curvature, already include the third dimension. (See also pp.
155 ff. of the present work.
43. M. Merleau-Ponty (1), pp. 394 and 304.
102 Lived Space
two-dimensional, flat figure-an observation that is completely
different from a genuine depth perception-then this appears to be
possible only on the basis of a primordial intuition of depth.
Meanwhile, one here overlooks a newly emerging problem. The
distinctiveness of depth does not lie only in that it is the "most
existential" of all dimensions (la plus existentielle), "locating" me
befare the things (par Jaquel je suis situe devant les choses); it no
less retains its exceptional position precisely because it can be
abolished on the pictorial surface while retaining its depth-a
surface is capable of presenting "space" that is given "in person"
and is nonetheless illusory (see pp. 124 ff.).
Berkeley's second presupposition maintains that depth perception
is nothing other than an associated lateral perception. Merleau-
Ponty counters this notion with the objection that an incorporation
of the view from the other side, and thus from the other corporeal
subject, would transfer the space univocally polarized by corporeity
into a two-sided, indifferent relationship. Space would lose its
center and depth would lose its sense. This objection follows as a
consequence of his fundamental position as ,an existentialist. He
gives special consideration to spatial depth becausl;) the "mineness"
of the world comes so obviously to presence in it: it is the depth of
a space whose center is in me, and only in me, and which can never
be a space for another. This is a philosophy which stresses the
exclusiveness of the "mineness" of Dasein and of the world. Even if
it does not discredit the entire domain of intersubjectivity, such a
philosophy truncates it considerably in its existential relevance. In
its conception of space, existentialism evidently has "no space" for
the aspect of the other. Thus it must necessarily Overlook what is
tenable in Berkeleyan thought, which, although veiled under an
untenable psychology of association, is nevertheless the underlying
basis.
If one lets this veil fall, then the core of the argument remains: The
position of the other, and thus the other himself, is incorporated. To
think that this incorporation would lead to a rejection, to a negation
of depth given "in person," is the mistake of Berkeley. He overlooks
the purely phenomenal state of affairs. In wandering about the thing,
in the discovery of the initially given depth as width, there is no
experience of deception, as though the three-dimensionally meant
thing were suddenly to appear as a surface. Rather, it is seen in a new
depth, and instead of abolishing depth the wandering continuously
reaffirms it. The ordering in accordance with surface and depth is a
pervasive structural moment of the space of intuition. That in
The Space of Intuition 103
wandering a thing will be grasped as three-dimensional, and that the
apprehension of the thing does not fall apart into a sum of exten-
sional moments, these make sense only because with the view of
each side the thing is "intuited" in its entire volume. The single
three-dimensional space is present in my depth dimensioned spate
of intuition, as it is in that of any other corporeal subject. It is in no
wise devalued simply because it is oriented toward me in my being
here and now; the space of intuition is deployed for the intuiting
consciousness in a multiplicity of possible aspects. It appears as a
"perspectiva!" space because it presents the hic et nunc view of the
one space for all corporeal subjects. What Leibniz has claimed for the
relationship between the monads and the world is valid for space:
they all reflect the universe in their own way; yet it lives in them all
as a whole.
Given the state of affairs-where the subject intuits space
perspectivally and conceives of himself as its center-the subject
ceases to be the subject only of his space. The subject possesses a
world that faces him, relative to his lived body, and a space
exclusively his own., Yet he knows the world, he knows space as
possible for others and theirs as possible for himself. This prefigures
a particular kind of substitutability still completely immersed in the
corporeal and stilllimited to the purely spatial meaning of "taking
the place" of another while remaining with oneself. Here the subject
in his corporeity already represents his intersubjectivity. It is impor-
tant to note that this is not something that can be attributed to the
individual subject nor can it be arbitrarily accepted or disregarded.
The depth of space is not the "most existential" of the dimensions
because the momentary location of my lived body provides a center
for an otherwise all-encompassing and positionless space, but be-
cause it affirms my participation in space as a corporeal subject-
with an ever different view and fullness.
The phenomenon of depth is closely related to that of perspectiv-
ity, and the observation of the latter may be newly illuminated and
complemented by what has just been said. The previous limitation of
our discussion to the things as being merely in front of and behind
one another may now be dropped. It may be noted that beside this
arder, they are also given in specific intersections, foreshortenings,
and "disruptions." Thus we must touch upon the central Husserlian
concept of spatial adumbration in the phenomenology of perception
to the extent that it implies the phenomena of form and place as
relative to the position of the observer.
The phenomenon of adumbration as a psychological, mathemati-
104 Lived Space
cal, or artistic problem must be left out of consideration. We shall
accept only the phenomenon of sensory intuition, from which all
subsequent problems originate. We shall consider only the subject
who is found in the center of perspectiva! space. He knows nothing
of linear or central perspective, nothing of geometry and wave
optics; he has the things for himself, and nothing further.
Being befare me, space offers itself in its depth; it is revealed in the
phenomenon of covering and being covered. The spatial thing is not
hidden solely because it is hidden by others but also because, in an
individual intuition, it is never completely discovered. The thing as
a whole is never given "in person." M y view is always an aspect and
encounters the thing from specific "sides." Each individual intuition
is essentially one-sided and is therefore inadequate in terms of what
it can accomplish in grasping an object.
44
The thing is accessible
all-sidedly only through motion, whether the motion is a succession
of one's own progressive movements ora spatial displacement of the
thing. Thus its all-sidedness first unfolds in successively going
through a series of positions.
The following problem then emerges: how can the successive
aspects give us "the" thing, which shows itself in them in different
modes of appearance and at the same time is present through them
all as something identical? How can an all-sidedness result from
many one-sidednesses? Can partial intuitions integrate themselves
into a total intuition of the thing? These questions become important
with regard to perspective: here the true form of the thing results not
merely from the sum of individual views but also from disrupted
side views; each successive one certainly complements the preced-
ing one, but also falsifies it and changes it anew.
Seen more exactly, such a formulation of the problem assumes a
specific standpoint and accordingly has a readily available point of
departure for its solution. In reality it does not permit the emergence,
through perspectiva! aspects, of an identical form of the thing, but
rather already presupposes it: the true form is one that is given
independently of position, unperspectivally, given in an intuition
that is no longer sensory intuition, no longer directly related to the
body. All changes experienced in the course of a series of aspects in
the intuition of a thing must be seen as disruptions of the true form
of the thing. That naive intuition does not progress from error to
error but can speak justifiably of grasping "the" thing means that the
44. Concerning the concept of inadequacy, see E. Husserl, Ideen I, 44
and 143.
The Space of Intuition
105
subject somehow has the true form of the thing. While traversing a
manifold of places, the subject is not continually required to verify
the thing in experiential "material" in each individual case, much
less to generate it.
This notion contains two main theses and accordingly requires
two kinds of explanations. (1) Each individual sensory intuition is
supported by a unifying moment; the contents of the individual
aspects are bound by the unity of sense of the "full thing of
intuition." (2) The thing as a unity of sense, retaining its identity
through the changing aspects, can be nothing other than the geomet-
rical physical thing. It can be counted as the true thing, thus the
sciences can determine the thing independently of position and can
conceive of itas it is "in itself."
The grounding of the first thesis usually has recourse toa specific
function of consciousness. Since Kant this has been introduced in
various ways under the rubric of synthesis in arder to account for the
constitution of unity. Using a house as an example, Kant has
presented the conception of an identical thing as an achievement of
"the power of imagination" in three syntheses.
45
Apprehension must
first collect the impressions into an "image," although the latter is in
no position to bring the object to presentation. This requires the
addition of reproduction, which connects the individual apprehen-
sions one with the other. Finally it requires recognition, which
orders the continua! successsion of reproduced images in a
superordinate synthesis, connecting them sensibly one with the
other, "in accordance with a rule," which allows the individual
"images" to become "sides" of an identical object. While this is
based on the strict Kantian distinction between matter and form on
the one hand and between forms of intuition and categories on the
other, the thing-conception is developed as something that is added
to the individual images. This clearly states that the sensibly intuited
thing is not given solely through sensibility but is partially achieved
by understanding. Thus apprehension and recognition are called a
synthesis in intuition, while reproduction is called a synthesis in
concept.
This conception requires a variation in Husserl's theory of contin-
uous syntheses.
46
In distinction to membered syntheses in which
discrete individual acts unify themselves into an articulated unity-
such that this unity is to be grasped as a new act of a higher arder
45. I. Kant, Vol. III (Transcendental Logic, second analogy of experience).
46. E. Husserl, LU, Investigation 6, p. 47; Ideen 1, 118.
106 Lived Space
founded in partial intentions-the continuous syntheses are distin-
guishable by a blending of various partial acts into one act; never-
theless, this act cannot be conceived as one of a higher level. On the
contrary, the unity of the object does not first emerge in a continuous
course of individual intuitions through a specific synthetic act, or
through a particular form of synthesis added to the partial concep-
tions; rather, the object is constituted as a "simple" unity without
recourse to new kinds of intentional acts.
In phenomenological description this state of affairs reveals that
each individual intuition contains the presence "for the time being"
of an aspect, and that this aspect is known as one among many
possible aspects. More precisely, this means that individual phe-
nomenological components contain implications of further aspects
in which the thing can appear. This possibility is "open" insofar as
no given aspect can claim preeminence with respect to its value in
presenting the thing; thus their series can proceed arbitrarily and
endlessly, though repetitions may occur. Yet the content of the series
is not completely indeterminate. In each individual intuition there
are anticipations of further intuitions. Their indeterminacy means
positively a determinability in the framework of a prescribed style.4
7
Thus in individual intuition we do not merely see forms but, rather,
"sides" of a thing which "belong" to the thing. In arder to observe it
all-sidedly, I turn and shift it and expect something from such
movements: the view of the "back"-side. A m ~ r e succession of
perceived forms ora serial arrangement of discrete, individual acts
would not induce identity. The total intuition, "thing," is that which
continuously directs further experiences of the thing and motivates
the movements relative to it.
We must add that this unity of intuition and movement is not only
a motivational unity but, above all, a functional unity. The unified
apprehension of a thing is not mediated fortuitously through
corporeity but has its basis in it: only a motile corporeity is able to
maintain, in the one-sidedness of a view, the ndices of further
aspects of the thing. These aspects become redeemable only in the
actual dynamics of corporeity. The sense of such ndices includes
the notion that they are viable solely for a corporeal subject.
The second thesis regarding the objective form of the thing as
geometrical-physical assumes a different signifigance. It is striking
that on the basis of this thesis, the perspectiva! adumbrations are
47. Concerning the concept of open possibilities, see EU, 21c; concern-
ing the meaning of anticipation, see Ideen 1, 44; CM, 22; EU, 8.
The Space of Intuition 107
conceivable only in privative modes of description: disruption,
truncation, deformation. While these terms are correct for formal
geometrical concepts, they are inappropriate for the actual states of
affairs. Thus in all the subtlety of his analyses of perspectivity,
Husserl could see in the adumbrations nothing more than the
subjective modes of appearance of the thing; and he quite unfortu-
nately chose the example of perspectiva! adumbrations to demon-
strate his distinction between the object of intuition and the concept
of intuition. In contrast to the physical thing as an object, the
perspectiva! adumbrations must be mere data of sensations belong-
ing to the really intrinsic [ree11] components of the perceptual
experience. Despite the ever greater emphasis on the distinction
between the adumbration and the adumbrated, Husserl failed to
recognize that even the former is not just an experience of conscious-
ness, but an adumbration of the experienced thing itself. Indeed, it is
an experienced mode of appearance of the thing and yet, as such, is
not a really intrinsic component of experience; otherwise that
component itself would have to be spatial. But such a notion is
obviously countersensical. That the perspectiva! adumbration is
continuously given as relative to the subject in his corporeal now
and here, that in addition the subject is intuitively co-conscious of it
as relative, does not change the fact that these adumbrations belong
to the thing itself-certainly not to the thing given in the categories
of objective science but as it is given in my corporeally-centered
space.
48
The phenomenon of depth makes it clear that this space is not
governed merely by factual affairs. 1 find myself in a situation with
things, and it is no accident that the alleged deformations resist
extension into depth. The perspective appears as a "disruption"
only to the extent that it is taken one-sidedly "perspectivally" on the
thing, such that the thing offers itself in only one "view"-thus
48. E. Husserl, Ideen I, 41. In the Husserlian phenomenological sense,
color and form-adumbrations would have to be attributed exclusively in the
perceptual noema; as really intrinsic components of experience they are not
to be found. That vision corresponds to a physiological process is not to be
denied, but in the phenomenoiogical attitude this process is not perceived
and thus it does not belong in an act phenomenology. Concerning a critique
of Husserl's one-sided noetic analysis of the phenomenon of adumbration,
see A. Gurwitsch. In contrast, in another place the noematic side gains its
full force with Husserl; see EU 16. (The special position of this work must
not be overlooked; see L. Landgrebe's foreword.)
108 Lived Space
disrupting itself and its sense. The perspective is not in the subject,
nor is it merely something in the intuited thing. It is a mode of the
insoluble and mutual relationship between a corporeal subject and
an object. It guarantees the subject his constant connection with the
world and makes his orientation in space possible.
These results suggest our agreement with sorne thinkers who
defend the significance of perspective against unjustified claims of
the natural sciences. Yet for us this significance is not sufficiently
explicated, and another question belonging to the problem of per-
spective is not answered: why is it that the intuition of a die is not
just a formal identity in the changes of perspectiva! aspects, but
precisely a perspective of a "cube with six squares"? Obviously the
mere identity of the thing is insufficient here. If the continuous
synthesis assures that the individual intuitions do not fragment into
a sum of individual perceptions, it does not provide the basis for the
fact that the form of the thing, as independent of position, is the form
of this and of no other thing.49
One cannot dismiss this problem as a mere conundrum with the
suggestion that the consciousness of the object already "has" the
form of a die with its six squares. Although this state of affairs cannot
be contested, the question still remains whether this mode of
intuition can be made comprehensible.
With the perspectiva! constellation of the die there are six distin-
guishable, frontally present views. Although they are still grasped as
"sides" of a die, each individual given is reducible to a singular
surface. The peculiarity of these six perspectives lies in their
disappearance. What does this mean for the thing of intuition and its
space, and what does it mean in particular for the "placement" of the
subject? Perspectiva! adumbration depends on spatial depth. The
exclusion of perspective means an exclusion of the depth of the
thing; as a fact, the die is given without depth in the view mentioned.
Thus there is a "perspectiva!" intuition of the thing in which depth
is dissolved and the thing is given as sheer extension. The die no
longer stands in a specific situation relative to the lived body, but
rather is located on a surface standing perpendicularly to depth,
which is annihilated. It has become pure extension.
49. We chose a die as a simple model, although the following reflections
are valid in principie for any form of thing. Since at this time there is no
discussion of geometry, the concept of "square" is meant only morpholog-
ically. (For the distinction between morphological and mathematical prop-
erties see pp. 184 of this work.)
The Space of Intuition 109
Where depth is dissolved, however, the subject ceases to be
corporeally in space-the die in the conceptual sense "with six
square surfaces" is in a space in which 1 cannot wander, in which 1
am no longer a lived body in a situation. It is a homogeneous space
without standpoint, a space whose subject is externa! to it. But if the
space of intuition is only the mode and manner in which the one
objective space is appropriated monadologically, then the subject's
intent to determine the "true" form of a thing-independently of
position, as identical for all subjects-becomes comprehensible.
Since a corporeal subject can only have an object through the mode
of individual perspectiva! intuitions, then he can validly and justi-
fiably address the perspectivally ordered world as the true world.
The problems of perspectivity do not arrange themselves according
to truth and falsity, but sol el y around the two poles of existence of
the subject as lived body and as thing-positing consciousness.
6. The Finitude of the Space of lntuition
The space of intuition is a finite manifold. The array befare me of
things one behind the other is not unlimited but reaches a determi-
nate end of the visual range. Although its boundary marks the line
where spatial things cease for my here, it is not a boundary in the
exact sense of the word; rather, it is phenomenally an indistinct and
unclear region of dissolving and dissolved contours, supported by
the co-consciousness of the "continuation" of space beyond them. It
is intuitively possessed as a horizon. The phenomenon of the
horizon is dual. The far horizon is to be distinguished from the near
horizon or inner-horizon [innenhorizont] constituting a limitation
within my space of intuition in the sense of a boundary of the
appropriation and differentiation of spatial structures.
The inner-horizon is determined by what in physiological optics is
called the optimal boundary of the capacity for resolutions; phenom-
enally, it is the indistinct region of limitation that still maintains the
presence of the spatial structures of a thing. It is thus a horizon of the
near-thing within the total visual expanse. The sizes of things are
located here; they diminish with increasing distance and reach the
lowest limit at the far horizon. At the far horizon of the thing the
interna! and externa! horizons coincide; they are dovetailed, as it
were. Yet in a certain sense both horizons are independent one from
the other; they vary, for example, in movement, and no firm rule of
their coordination can be established. This is most clear in the
various possibilities of displacement of the horizon.
110 Lived Space
To begin with, each approach toward the things means displace-
ment of the inner-horizon. With the progressive clarification and
deciphering of the spatial structures of the object that occurs with
each approach, there results a continuous displacement of the
dissolving boundaries. Yet there is no necessary connection between
this receding of the near horizon and a displacement of the far
horizon. The displacement of the inner-horizon is not perceived as
though it shifted toward the "outside" but rather as a receding
toward the "inside"; the movement will not be experienced as a
protrusion but more precisely as an intrusion. (We often speak of
"deeper" intrusion into the "inside" of things when our basic
concern is merely to grasp the spatial structures of the surface more
precisely, in the course of which the inner structures frequently
come into view, as for example in microscopy.)
The greater stability of the far horizon, in contrast to the inner-
horizon, appears when one observes the forms of approaching that
are equivalent to one's own movement of nearing. With respect to
the achievement of clarity, the drawing near of the lived body is
equivalent to the things "drawing near"; both equally achieve a
displacement of the inner-horizon. The latter can be obtained even if
the lived body is stationary, while the far horizon is displaceable
only when one actually moves oneself. One's own movement
thereby already assumes special status; thus we must consider such
movement, as a specific characteristic of corporeity, with regard to
the principie of relativity.
It is important to note that there is no such relativity between
corporeity and the thing. While the movement of the thing can in
principie be compensated for by the movement of one's own lived
body, this equivalence of both courses of movements can be grasped
only with respect to the thing itself, not with respect to the
movement as such. The movement of corporeity is always absolute
in the sense that its specific inner experience (as kinaesthesis,
corporeal sensation, etc.) reveals the moving body immediately. A
relativity of movement is possible only among moving things. 1t is
here that the conception of the "assumed" object at rest attains its
realization; in the space of intuition, such an assumption remains
constantly bound to the specific conceptual senses of the natural
context of things evident in the common example of a departing train
and the "stationary" station. This assumption can first attain full
indepe11dence and the form of a principie in a space that is no longer
ontologically relative to a lived body.
The extension of the inner-horizon within the space of intuition is
The Space of Intuition 111
nothing other than what was previously called the pervasive
structuration of space. Within the space of action there is already an
essential function of implements, which accomplishes progressive
differentiation and leads to an ever finer differentiation of the there
and the yonder. While in this space the articulation reaches only up
to the region, in the space of intuition it can be extended further by
apparatus accomplishing further decomposition and clarification up
to the phenomenal point. Even the latter is not a fixed size; as a
"point," its value is relative to the means used for intuition. Within
these means, what is ultimately given cannot be articulated any
further; rather it vanishes abruptly. Yet precision technology regards
the attained level of research and instrumentation of the phenome-
nal point as itself merely a region again capable of further decom-
position. Its stepwise advance into further levels of precision is
motivated by the claim of progressive condensation of the topolog-
ical nets all the way to the "ideal," i.e., really unreachable, limit case
of a mathematical point.
Nonetheless it must not be overlooked that in the use of these
implements there is a metabasis eis allo genos of intuition in the
truest sense of the word. The apparatus of "magnification," seen
from the side of the phenomenon, is indeed named appropriately. It
aids the corporeal functions of vision and assists toward magnified
clarity, exactness, and differentiation. Y et its signifigance and
achievement is not exhausted in being a mere analogue of sensory
intuition. Subsequent reflection must clarify the meaning of a
replacement of the visual ray by a light ray. These two concepts must
be strictly distinguished. The latter is used in the framework of
science as a physical concept having its special position in a
hypothetical-deductive system whose concepts can definitively be
alloted various meanings. This is not the place to discuss these
issues since even the concept of the visual ray requires clarification.
After all, seeing as a "raying" event in lived space does not lead to
an intuitive fulfillment of its meaning and cannot be phenomeno-
logically brought to light. Nowhere are visual rays encountered, but
always things in space (see further pp. 248 ff.).
New determinations are to be grasped at the far horizon. We must
recall the elementary orientational directions and the corporeal
movements corresponding to them. While the being who is oriented
toward the far horizon is at rest, it captures and encompasses the
alignment of things in the distance in one glance. In the limit case,
above, below, right, and left shrink into a phenomenal point. The far
horizon is that place or line at which the elementary extensions of
112 Lived Space
things vanish. It is especially the leveling of the opposition between
above and below that makes the horizon a "horizontal" and intrin-
sically limitless, although finite, line for the disappearance of things.
Depth and perspective vanish with it; the most distant things are
seen as flat and when in movement are no longer grasped as moving.
Heidegger has pointed out that the finitude of the space of
intuition is related to the finitude of Dasein. 5 The limitation of this
space means at the same time limitation of one's own being as living
and corporeal; the finite space is a space of a being who is and ends
in time.
This corresponds to our own point of departure. The signifigance
of the horizon-phenomenon is grasped primarily in the intuition of
motion. In the nearing and distancing of things the boundaries of my
space of intuition reveal themselves not only as what they are
ontically, but also what they mean ontologically for m y being as a
corporeal being. In the phenomenon of the horizon, spatial and
temporal determinations pervade one another in a remarkable way.
Horizon is the transition of things in intuition from the not-yet to the
just-already, from the still-there to the no-longer. One may look atan
auto speeding away and vanishing at the horizon-at the moment of
"vanishing," is it a spatial nothing or a temporal no-longer? As it
"emerges" at the horizon, is it suddenly grasped there or does it
already appear? Even a more precise analysis would show that both
determinations flow inseparably into each other, that here temporal
and spatial moments are simply not separable.
The horizonality of the space of intuition leads to the recognition
of its time-space character, and its limitation points to an original
unity of space and time, although quite unnoticeably and almost
imperceptibly. In intuiting it is the orientation toward the location of
what is emerging that remains in the foreground. This is in confor-
mity with our previous observations that the space-time as time-
space is experienced to the extent that the experiencing subject is not
aware of time itself explicitly. In the intuitive mode of being, the
subject is in principie aware of time as objectively as of space. He
apprehends the horizon as horizon of space only on the basis of a
separation that has already taken place. Correlatively, he knows how
to determine his finitude as being "in" time only through limitation
in an objectively known time. The separation into two objective
"forms of intuition" is itself constitutive for the horizon as time-
space phenomenon.
50. M. Heidegger, 23.
The Space of Intuition 113
Yet the far horizon is a boundary of space only in an inauthentic
sense. What is decisive is that the experience of a boundary contans
a consciousness of the "continuation" of space. The emergence and
disappearance of things at the horizon is not perceived as destruc-
tion and re-creation, but as coming from and disappearing into the
"beyond." It is precisely this limitation that reveals the space of
intuition notas something closed u pon itself but as a section of "the"
space: as a lived space relative to the lived body of a subject who has
itas lived space, in this arder andarticulation, only insofar as he as
corporeal subject has already transcended it.
The horizonality of lived space, seen by Heidegger as an exis-
tential of Dasein in its corporeal finitude, is an admission of a mode
of being of Dasein in which Dasein is beyond its finitude. Dasein not
only has a corporeally-centered space as its own, but rather is in
addition objectively aware of the one space as space for all corporeal
subjects.
7. The Other in M y S pace of Intuition. Questions of
Homogenization
In the space of intuition things are in front of me and for me; it is
my space. Through its depth and horizonality it testifies that it
belongs tome. Each of its things is a counter-pole to my corporeity
here and now; it is a thing with me in a given situation.
But not only things reside in it. My discussion up to now, resting
on a solipsistic corporeal subject, was an abstraction attainable only
retrospectively. After all, there are "things" in it that I can intuit as
having, like all other things, their location in my space, and yet
which are fundamentally distinct from all things. I know directly
that they too are a lived body. And lived body means a corporeity
like me, and thus a center of its space of intuition distinct from mine
and insurmountably el o sed to me in m y own "here."
How is it that the multitude of spaces of intuition results in the
conception of one space known to us in common? How do the
infinitely many aspects allow themselves to be integrated in the
intuition of "the" space?
There is no calculus that could achieve this, nor need there be one.
Rather, the questions just posed must be unmasked as absurd and the
preceding deliberations as inconsistent.
They started with a multitude of perceivable lived bodies and their
spaces and inquired into the emergence of "the" intuition of space.
I conceived of the other as lived body, and thereby primarily in
114 Lived Space
purely external spatial relationship tome; if my relationship to him
were based only on sensory perception of his physical body, then the
questions posed could be significan t. lt is different if 1 grasp the other
not merely as a physical body but as the lived body of "another" who
is also an "ego". Such knowledge is more originary and precedes
mere sensory perception of his physical body. lndeed, 1 never know
the other without corporeity, since he is corporeity. Yet to perceive
what is "there" as corporeity 1 must already know what makes it into
corporeity-in arder to be another being for me, it must also be an
ego.
This "also" is not that of a spatial thing external tome. The other
as an ego is never to be found in my visual field, just as 1 am not to
be found in this way for him. lnsofar as we both know ourselves as
egos, we stand in a relationship that is not merely non-spatial, but
first allows me to say something about it and to raise questions about
its spatiality. As an ego, 1 am completely determined through the
other, as he is determined through me. Asan ego, 1 am not befare him
or subsequent to him; we are both egos, since the one is only an ego
for the other.
Thus the question concerning the one space acquires another
accent. lt is not to be asked how the infinite multitude of distinct
spaces of intuition results in the one, identical space for all lived.
bodies; rather, what is to be discussed is how the singular space of
intuition is constituted in whose center there is a being who is not
only a motile corporeal being, but also specifically one who tran-
scends his corporeity, and whose ego turns out to be conditioned by
other egos.
With the emergence of the other there also appears a difference
that separates the space of intuition from visual space. We observe a
thing-the other from his location and 1 from mine. He points to it:
this corporeal motion means more for me than a mere outstretching
of one of his bodily members. With it I immediately understand the
situation of the other, his orientation toward something grasped by
him in a purely objective confrontation. He further points in arder to
"direct" meto something. Thus in his corporeal posture he already
represents not only his being with the things, but his being with me.
A solipsistic subject does not point; pointing is understandable only
with a corporeity that is a corporeity for others, and thus is capable
of being the corporeity of anego. What is crucial is that he can point
out "the" thing to me- a senseless undertaking if he meant only the
visual thing, which is different for him than for me. The aim of
pointing has already surpassed the latter toward the wholeness of
The Space of Intuition 115
the thing, grasping it all-sidedly and knowing that I too intuit it in
the same manner. Our agreement does not aim at the visual thing but
at the intuited thing in its meant all-sidedness.
The identity of our being, as a consciousness of an object, is
rediscovered at the thing insofar as we are oriented toward it-not in
the manner of vision but of intuition. This is a categorical intuition
inexplicable solely in terms of corporeal functions; thus it can be
maintained that in my space of intuition, as well as in that of all
others, there is an intersubjective moment graspable on the thing as
an identity. But the thing is not now meant as identical in the change
of its perspectiva! adumbrations, but as one for all subjects like
myself capable of categorical intuition. Naturally this has nothing to
do with two identities. Our previous observations touched only
u pon a single subject, as if to imply that the others were not there, as
if they did not belong to my components of intuition-and yet
although they were actually not present, they were nevertheless
co-posited. (The "real solus ipse would be completely incapable of
such co-positing; he could not conceive of an alien ego-corporeity in
the full sense. Even if he were conscious of his own corporeal
functions, he would have no awareness that his corporeity is for
others, and thus he would inappropriately assume the title "ipse.")
Insofar as I recognize the other corporeity as also being an ego, 1
simultaneously allot ita place in my space, a place that is not merely
a there but also a here. The here thus loses its uniqueness; my here
turns out to be a there for the others. M y space of intuition is indeed
mine by virtue of my corporeal being, but for me it is any space by
virtue of my consciousness, and thus a space in which it is possible
to have a harmonious experiential context with others.
The guarantee of such harmony is already present in the corporeal.
The here of the other is not simply his here; at any moment it can
become my here by virtue of my capacity for movement. Phenome-
nally speaking, my movement contains the loss of a center; the
movement into spatial depth, revealing in the succession of things
their contiguity, announces the homogenization and thus its loss of
limits. Its horizon is a horizon relative to my immobile corporeity-
its retreat is my advance, and thus the "continuation" of space is
nothing other than my movement.
Yet my movement creates and recreates a new center and a new
depth. In this sense, the intersubjective space as a whole cannot be
unfolded, since the factual movement extends only across a deter-
minate part of space; due to the finitude of the corporeal being, an
endless movement is impossible.
116 Lived Space
The problem emergng here has already been noted by Scheler.
Scheler, too, seeks to explcate space as a space of movement. For
him space is not at all an object of any intuition-thus Scheler
criticizes Kant for grasping space as a purely perceptual formation-
but solely a vital achievement of a motile body, of a being of drives
"having nothing to do with reason and intersubjectivity." Scheler
poses the following question: how can the limited possibilities of
movement of a finite corporeal being account for the understanding
of space as infinite? The answer is that space as such remains merely
a "possibility" of a capacity for self-movement, while its actuality is
descriptively introduced by a consciousness capable of conceptual-
izing space. Thus the "immense paradox" of the one, substantial,
empty space comes to the fore; a space that the subject, on the basis
of dissatisfied dynamic drives and his unlived and unlivable possi-
bilities of movement, "projects outward. "
51
With this point of departure Scheler has undoubtedly touched a
nerve of the entire problem of space. He has discovered a defect in
the traditional theory of space. The latter was accessible only to a
philosophy oriented from a pure consciousness. Space is space of a
being who is irrevocably its own lived body-it is a space of
movement in two senses: a space for, as well as constituted through
the movement of a corporeal being; this was shown in the preceding
in vestigation.
Nevertheless, Scheler's solution remains unsatisfactory in various
respects. First, it must be asked whether the unlived, constantly
present overabundance of drives, the held-back vital powers, con-
stituting, according to Scheler, the difference between humans and
animals, is adequate for the comprehension of infinite space. It
seems that the ego-corporeity must contain an entirely different and
moved structure of drives: they must be entirely different from
merely vital activities, and yet they must be understandable in terms
of the ego-centric, the corporeal subject. His being as a conscious
being must at least pro vide the conditio sine qua non for that specific
experience of movement allowing us to seek the foundations for the
consciousness of endless space not in the mere fact of my capacity
for self movement, but in a consciousness of this ability. That such
a consciousness should be liable to succumbing to a deception,
confusing a mere "possibility" for "actality," constitutes a thesis to
which Scheler must revert without undertaking any justification for
such a thesis-which would have required the precise clarification
51. M. Scheler (3), pp. 295ff.
The Space of Intuition
117
of the meaning assigned to the concepts of possibility and actuality
in the sense meant here.5
2
Secondly, it remains completely incomprehensible how a projec-
tion of space, conceived by Scheler as a purely vital activity, ought
to bring about an identical space-consciousness for all ego-
corporeities. It is also difficult to discover how a qualitatively singular
experience of overabundance could become conceivable for con-
sciousness as a quantifiable and mathematizable emptiness.
Finally, Scheler overlooks the problem of the topological structure
of this space. The presupposition of the open-infinite manifold,
accepted by Scheler without question despite the fact that it is most
problematic, does not follow convincingly or univocally from his
own basic conception. If the reservation mentioned above were
removed, it would be possible to agree to an endless-closed structure
of space curving back upon itself, particularly when Scheler repeat-
edly characterizes corporeal movement as periodic and rhythmic.
Yet the space constituting consciousness cannot simply, arbitrarily
decide to represent space in one or the other manner. On the other
hand, consciousness cannot perceive the one open-endless space as
necessary: indeed, it is precisely here that it cannot be maintained as
necessary, since it cannot be presented through any specific complex
of phenomena. It is impossible to derive an intentional content of the
historicity of consciousness from the history-less constants of vital
structures.
In any case, it must be pointed out that we hardly have sufficient
knowledge about the history of the unreflective experience of space.
What unfolds as a history of the problem of space is the history of
theories of space, but not of space consciousness itself. From the
beginning, these theories were by no means articulated completely
in thoroughly phenomenological description; they remained pre-
dominantly speculative orientations toward cosmic-metaphysical
questions. In turn, however, the theories were the conceptual
sedimentation of a spatial conception, which, although not created
directly from the natural intuition of space, was not in an open
contradiction with it. Thus the theories mediately present the
history of space intuition.
53
52. In addition see pp. 157ff. of the present work.
53. Phenornenology has not yet recognized the task of investigating the
philosophical past as historically problernatic in terms of its own rnethod-
ological style. We shall return to this question (pp. 157 ff.).
118 Lived Space
8. Open Questions.
The problems just alluded to lead beyond the region of what can
be shown descriptively into the domain of the metaphysics of space.
To pursue them would be the task of broader discussion. At this
juncture only a few of the remaining open questions ought to be
raised. We cannot simply blur over the specific difficulties that have
appeared in the last discussion.
They discovered the subject to be in a double attitude toward the
space of intuition: on the one hand, the corporeal comportment "in"
it, and on the other, the theoretical attitude intuiting "the" space
objectively.
While the space of intuition turned out to be co-determined by,
and not detachable from, objective space, the converse was not true.
Thus the question remains whether the objective conception of
space, which we assumed to be contingent datum of consciousness
as such, is capable of being traced further back, or whether it must be
taken asan ultimate factor, asan irreducible structural characteristic
of "the" space-consciousness in its own specific character. Viewed
in terms of space, we must ask whether space, as a "pure form of
intuition," is to be related to the space of movement, or whether its
basis lies rather in the conception of space as space of movement.
To recognize space as both, to reduce it to an unconnected
"both/and" of the peaceful coexistence of two distinct "aspects"
would evidently run counter to the meaning of a philosophical
theory of space. But to allow one of the two to be solely valid would
be equal to the truncation of the perceivable phenomena discovered
in the framework of our analysis. The conception of space as a mere
form of intuition is powerless to explain the interrelationship of
corporeal movement with space, specifically with regard to the
plethora of individual phenomena present in the observation of
various corporeal modes of comportment. In contrast, Scheler is
exposed to the danger of a reflexive circle. His projection theory
attempts to derive the objective consciousness of space from a
non-conscious structure of drives, which, in arder to be appropriate
for the constitution of "the" space, must have assumed non-vital,
intersubjective moments of consciousness.
These doubts cannot be directed against our investigation, since it
did not undertake any attempt to deduce the objective space-
consciousness from previous achievements of the corporeal subject.
Nevertheless, it must defend itself against another critique. Again
and again our attention has been drawn to the fact that the investi-
The Space of Intuition 119
gation has traced the path of a subject through differing spaces,
which the subject has "always already" traversed. The manner in
which space is possessed at a given level of his comportment is
decisively co-determined by his objective consciousness of space.
Thus our investigation must take more seriously the charge that the
conception of space as a space of movement of a corporeal being has
been surreptitiously attained after all, that in reality space for this
being remains nothing other than a form of intuition of conscious-
ness.
Our investigation has not been able to bring together these two
factors. The answer to this question lies beyond phenomenological
demonstration. After briefly touching upon the modally distinct
sensory spaces in the following chapter, the next section deliberately
abandons the phenomenological domain in the narrower sense
without negating it. The immediate aim is to devise a theoretical
foundation capable of elucidating the phenomena dealt with until
now.
Chapter Four
Modally Distinct Sensory Spaces
1. Visual Space
It must be emphasized at the outset that the sensory spaces to be
considered here will not be developed as fully as was the space of
intuition. The latter turned out to be relative to a specific, self-
sufficient mode of being of the intuiting subject. In contrast, the
modally distinct sensory spaces are present only as nonself-
sufficient. In arder to reveal their structure, it is necessary to assume
a particular abstractive view. The subsequent limitation to visual
and tactile spaces does not constitute a prejudgement concerning the
existence or nonexistence of other sensory spaces-such as an audial
space-but rather serves the explicit purpose of clarifying the space
of intuition. The two designated spaces relate to it as two abstract,
nonself-sufficient moments of a concrete whole.
Yet in no wise does this mean a construction of this whole from its
parts or its breaking up into parts. Rather, we are solely concerned
with abstractively seizing upon all those essential traits of both
sensory modalities, which as a sum never comprise sensory intu-
ition, although they can shed sorne light on the latter.
The separation of visual space from the space of intuition encoun-
ters sorne difficulties. Their difference does not lie in differing
objective data, but only in their differing apprehensional sense. As a
thing of intuition, the thing is accepted as "full." Though it is
constantly perceived one-sidedly in adumbrations, it includes ev-
erything co-given with it, and is grasped as a thing. It is only in this
manner that it can correlate to a possible world of the subject. In
concrete intuition, the merely perceived is already and always
surpassed toward the whole of the object as a specific unity of sense.
The abstraction that must be performed in arder to move from the
intuited thing to the visual thing requires the bracketing of the
co-perception and the reduction of the complete thing of intuition to
what is visible "on" it. It is to be maintained that the visible is meant
120
Modally Distinct Sensory Spaces 121
qua "visible on the thing." Strictly speaking, the co-perception of the
thing is not completely annulled but only consciously separated as
such from what is genuinely perceived.
54
An apprehended residuum
of the thing as a whole remains present and available in the view of
an object at rest. The movement of the thing elicits the apprehension
of the changing visual contents as "sides of the thing." Indeed, the
apprehension of the "side of a thing" is pushed into the background
by the intuitive attitude and its new visual content. Thus although
this apprehension does not completely break down, it is present only
marginally. While the apprehension of the moving thing of intuition
offers merely the other side of the same object, visually speaking
there is a clear consciousness of the complete otherness of the visual
object. What remains of the object and its space in pure vision after
thus bracketing the co-perceived?
First, there remain sorne of the characteristics analyzed in the
space of intuition, e.g., its peculiarity as a phenomenal manifold of
points-which have basically been revealed with the phenomenon
of the visual thing (pp. 85ff.)-as well as those properties that are
related to the finitud e of the space of intuition. But they clearly show
that the space of intuition is not exhausted in being mere visual
space. In any case, the limit of the space of intuition is the same as
that of the visual space; the domain of the intuited things extends no
further than that of the visual things. Yet in intuition it did not
genuinely appear as a limit, but as a horizon in the more precise
sense that the latter is perceived as the region of the not-yet and the
no-longer beyond which space "continues." This co-consciousness
of the continuation motivates the anticipation of the new experien-
tial context and determines the sectional character of the space of
intuition.
It is otherwise in visual space. If one remains strictly with what is
given in vision alone, then nothing remains of the continuity of
space. Visual space does not have a horizon but a limit, a region of
the literal disappearance of the thing. The visual thing as such is
either given or not given "in person"; it does not, however, have the
not-yet and the no-longer. The same holds for any perspectiva!
adumbrations. Strictly speaking this concept too belongs to the space
of intuition; the visual thing does not hove adumbrations, rather it is
adumbration. That is why with every position and after every
movement the visual thing is a different one; it appears differently
54. The complete annulment of co-operation would lead to the annihila-
tion of spatiality. Compare the analyses of the visual field, pp. 124 ff.
122 Lived Space
relative to the position of the observer. Thus all discussions of the
phenomena appropriate to the clarification of the monadological
traits of the space of intuition were, strictly speaking, oriented toward
things as visual things.
That the distinction between the two "spaces" is slight makes it
understandable that they are usually identified. The lack of their
differentiation constitutes the ground for the controversy among the
opponents dealing with the "euclidean nature" of the space of
intuition; the numbers in the opposing camps as well as their
apparent power of proofs are somewhat equal. The concept of
"euclidean nature" is associated with the existence of parallels in
the intuited (morphological) sense of the orientation in the same
direction, i.e., non-intersection of two straight lines, as well as the
constancy of an object through specific movements, rotations, trans-
lations, mirrorings; this means that from a space determinable as
euclidean one expects the same intuitive fulfillment of those require-
ments demanded of euclidean geometry as a special theory of
invariants. Obviously, for the visual space these requirements are
not fulfilled; it does not have parallels in all directions, and the
shape of the thing and its size are functions of the location.ss In
contrast, a euclidean nature can be meaningfully attributed to the
space of intuition insofar as the rift between what is itself given and
what is co-given vanishes and the objects are "intuited" all-sidedly
in any motion. Parallelism is also guaranteed here. The controversia!
question of railroad tracks stretching into depth is answered: they
are visual things, two strips approaching one another, while as
objects of intuition, they "run" parallel-they stretch further, be-
yond the horizon, a situation which is already implied in the
conception of "railroad tracks."
Preliminarily, this is the sale justification for the claim that
despite its centricity, the space of intuition is euclidean-more
precisely, that the same metrics can be applied to it as to the
homogeneous space of objects. It is of course necessary to show that
this is the euclidean metrics. Here we wish only to suggest that the
solution to the problem depends on a precise formulation of the
question andona clarified terminology.
55. The change of size of visual things is not constant; in a specific region
of nearness, there appears the remarkable phenomenon of size constancy.
Thus visual space becomes nonhomogeneous in a dual sense: it is not only
centered on the point singled out as my here, but it is also articulated
concentrically into various steps of distinctive distances.
Modally Distinct Sensory Spaces 123
A more problematic question is whether the reduction of the full
thing of intuition to its purely visual givenness accomplished here
does not obliterate space altogether and destroy its depth which is
implied in the things being in front of and behind one another. Such
would be the case if this ordering of the things were not retained for
the visual objects. If one were to orient oneself toward a specific
thing-constellation and perform the reduction under discussion,
and if sorne change were to occur in the region of the visual thing,
one would find that despite the appearance of a formal deformation
of the visual thing, one would still apprehend the change as a
movement and not merely as a transformation of forms on a surface.
The perception of movement is an accomplishment that cannot be
carried out solely through the visual function. The sensory basis of
constitution of something seen as moving cannot be provided solely
through vision; rather, it requires the complementary activity of
touch.
The functional unity of vision and touch has been stressed for a
long time. The perception of a thing was already conceived by Locke
and Berkeley as originating from the associative connection of visual
and tactile data. Berkeley in particular repeatedly stresses the strict
distinction between the visual and the tactile function as well as the
fundamental role of the tactile sense for the construction of the
perceptual world. For him the visual is only a sign for the tactile, and
touch alone "suggests" to the rest of the senses their objectivity.
56
Berkeley's influence is found in Husserl's theory of aesthetic syn-
thesis. Underlying the genuine thing-giving synthesis, it has,
according to him, the function of uniting the tactile with the visual
data into an appearing schema.
57
Just like Berkeley, Husserl also
maintains a strict separation between the visual and the tactile thing,
to such an extent that he attributes self-sufficiency to the purely
visual thing (spatial phantom) rather than to the tactile thing.
Yet it must be noted that even where the discussion focuses purely
on the visual thing, no abstraction can be made from its materiality.
It is a mistake to think that what is not and cannot be given in "pure"
vision must completely cease as a datum on the visual thing. Insofar .
as a thing is conceived in any sense as thing-like, its materiality is eo
ipso co-given phenomenally in a "pure" visual datum such as color.
After all, the color is not merely a color medium externally attached
to the thing; rather, in its other mode of appearance-shining,
56. G. Berkeley, 117-19, 125.
57. E. Husserl, Ideen II, 10.
124 Lived Space
glittering, shimmering, transparent, dull-it betrays something of the
material structure of the thing. But the materiality of the object as
such is founded sensibly in touch.sa
Furthermore, the activity of the tactile sense must already be
presupposed if the vision of movement is to be possible at all. The
criterion for the vision of movement les in that the moved can in
principie be "followed by the eyes." Yet such following is insuffi-
ciently understood if it is apure movement of the glance; this as such
must also be present to the body, i.e., it must be experienced "from
within" as oculomotoric, and it must be able to be made evident as
motoric. This typical inner experience of a movement is, however,
necessarily accompanied by tactile sensations. The same is true of
more extensive movement of things, which can be compensated for
through the movement of the body, thus from within, kinaestheti-
cally. The indicator for a movement is therefore ultimately the rest or
movement of the lived body, or its limbs, which can only be
experienced as the rest or movement of a "body" capable of tactility.
A corporeity capable of vision but not of tactility would know no
movement, not even its own, nor would it be cognizant of a visual
space; the latter remains ontologically relative to a corporeal unity of
visual and tactile functions.
2. The Visual Field
It is not contradictory that through a specific change of appercep-
tion the spatial manifold of visual things can be transformed into a
field conceived as purely extensive. The moving visual things
rigidify into a merely superficial arrangement of colors and forms,
and their depth is reinterpreted as a mere next-to-one-another and
one-upon-the-other of figures with specific intersections. A being-
one-behind-the-other, still graspable in visual space by virtue of
specific displacements, merges here into a purely extensive, merely
figural change.
The bracketing of perceived movement abolishes all residuum of
"thing" -apprehension and thus also the conception of emptiness as
"leeway" for the movement of things. Surface confronts surface,
color touches color without any intervals; the doubly reduced and
thus disciplined vision ultimately no longer perceives color as a
mode of appearance "on" the visual thing, but merely as extension.
Gleam, shimmer, etc., "on the thing" are transformed into an
extended "spot" of color.
58. W. Schapp, pp. 19 ff., for additional aspects see also D. Katz (2).
Modally Distinct Sensory Spaces
125
This level of abstractive vision is required for a pictorial
presentation of the painterly spatial perspective, since the latter is
precisely a depiction of three-dimensional space on the pictorial
surface. Seen strictly phenomenologically, this does not have to do
with a presentation of space but with a transposition of the
abstractively attained visual field onto the painted surface. It is no
accident that the full mastery of the painting technique of the
"intervals," i.e., of the empty space in between things, appears quite
late in the historical evolution of painting. It not only requires the
"eye of the artist," but also a vision in the sense we have sketched
out here. In addition, it is necessary to have an aesthetic attitude
that is first attained from natural intuition through a specific
reductive undertaking. Thus from the painted figures on the
pictorial surface, the three-dimensionally apperceived object imme-
diately "leaps to the eyes" of an impartial observer, as long as the
laws of linear and painterly perspective are maintained within the
broadest limits. In any case, it is also possible here to remain in the
necessary attitude toward the pictorial presentation itself or to
choose between or shift among various possible attitudes, and thus
to call forth the impressions in accordance with the requirements of
the so-called originary experience.
A more detailed exposition of this problem of presentation of
space would lead us too far astray from the true aim of our in-
vestigation.59 It must be kept in mind that visual space phe-
nomenologically precedes the visual field and that the former in
its turn is conceivable only as an abstractive moment of the space of
intuition. Since it is not a "purely" visual space, but is constituted
through vision and tactility, it already appears as a relatively
complex structure. A further step of abstraction leads to the visual
field, which is difficult to describe phenomenologically. It is
occasionally grasped when the possibilities of coincidence of visual
and tactile data come into view. Y et a separation of the originary
unity of coincidence of the visual and tactile thing is phenomeno-
logically problematic. Even pure surfaces require the tactile
function in arder to appear visually as surfaces. Where the surface is
not really touched, but in a transferred sense is "touched" with the
eye, the tactile sense is still required, just as the kinaesthesis of a
lived body capable of tactility must be present in the perception of
movement. Thus a further question appears: whether and to what
extent touch in its turn is to be conceived as a function that presents
59. On the pictorial presentation of space, see E. Stroker.
126 Lived Space
objects and space, and particularly whether it is constitutive of an
independent tactile space.
3. The Problem of Tactile Space
It is an old insight that the tactile sense plays a constitutive role in
the cognition of objects. Kant points out that it is the sale directly
perceiving sense, and thus the most important and instructive.
Subsequently, Palagyi emphasizes its fundamental role for all per-
ceptual knowledge. In the same sense Katz, who is the first to offer
a precise psychological analysis of touch, ascribes to it a decisive
character of reality and primacy for all sensory knowledge.
6
o
This remarkable activity of the tactile sense still does not guaran-
tee the existence of an independent tactile space. This question can
be decided only through an exact investigation of the objectivity
given in tactility. Precise analyses must be undertaken concerning
the manner and mode in which tactility and its presumed objectivity
are given with its respect to function as a sense and its unique
position among the rest of the senses.
Katz had already recognized that tactile phenomena are bipolar
insofar as they result from a subjective ( corporeally-related) and an
objective (oriented toward the objective characteristics of things)
component. As a result, the terms tactile perception as well as tactile
sensation have a justifiable meaning. In the scale of sensory func-
tions, the tactile sense assumes an intermediary position between
sensory impressions of a state or condition and sensory-perceptions
yielding objectivity.
In accordance with this bipolarity, our investigation will have two
directions. On the one hand, it will survey the mani:qfold of
locations of a tactile there with regard to specific spatial structures,
and on the other, it will reveal the specific manner and mode in
which corporeity is given tactility. The latter in its turn is twofold: in
tactile sensation and in kinaesthesis. This complex problem consti-
tutes the specific difficulties of the following investigation. The
problem inheres in the essence of the state of affairs. Although our
observations are concerned primarily with the question of tactility
and the possible discovery of tactile space, these distinct aspects
60. I. Kant, Vol. VII (Anthropology, Part One, 17); M. Palgyi (2), 3rd
Lecture; D. Katz (1), pp. 255 ff.; additions to D. Katz in R. Pauli's work.
Modally Distinct Sensory Spaces
127
cannot be strictly separated in arder to show their subsequent
intersections. B1
The existence of a tactile space long remained indubitable. Yet it
is characteristic of its particular problems that its first thematic
investigation also led to its denial.
The first answer to the question posed here did not come from the
area of philosophy, but from that of pathology. In the investigation of
brain damage, the neurologists Gelb and Goldstein were led to the
view that "spatial characteristics do not appear through the qualities
mediated by the tactile sense .... Spatiality enters the tactile domain
only through visual representations, i.e., there is only one genuine
visual space."
62
Their investgation, which gained a great deal of
attention in neuropathology, became a catalyst for the revision of
basic principies acquired in other ways. For example, using their
research Schilder placed his conception of a "body schema" on a
new basis. Their investigations are worthy of mention on two
grounds. First, the complex of problems relating to tactile space was
confronted for the first time; second, they share the fate of many
initial works: with the development of a new problem, they also
reveal their own felt lack of a solution. Thus in the work mentioned,
it is not clear what is to be really understood by "tactile space,"
specifically when it is taken as something that is not on hand. What
Goldstein and Gelb bring out is the inability of a patient to localize
tactile impressions of his own body. Yet nothing is thereby decided
concerning a tactile space. Above all, the method used by the two
researchers to attain their results must appear dubious. If their work
did not propase anything more than an attempt to describe and
explain mere behavior, remaining an internal concern of neurology,
then it would be justified, and philosophy would have no right to
assume a critical stance to it. But their case is expressly different:
from individual analyses of extremely complex pathological find-
ings, they make a claim that purports to be valid for the tactile
phenomenon as such, and thus pretends to be an essential proposi-
tion in a phenomenological sense. How little both investigators are
clear on their own methodological stance already follows from the
view, for example, that the execution of arbitrary movements by the
patient is similar to that of a normal child. Besides, the exposition
moves in various circular conclusions. The most significant one is
61. The next section will investigate more closely the aspect of touch as
sensation and the constitution of the givenness of one's own body in touch.
62. K. Goldstein and A. Gelb (2), p. 73.
128 Lived Space
found in the localization hypothesis: the inability of the patient to
localize tactile impressions on his own body is reduced to the lack of
optical perceptual capacity, and from there it is concluded that
spatiality does not pertain to the manifold of tactile data themselves,
but rather that spatiality is constituted solely by the visual sense; yet
it is assumed in turn that the latter is mediated by the kinaesthetic
residua awakened through the tactile impressions.
Concerning the necessity and universality of the optical capacity
of representation for tactile localization, it has been conclusively
shown by experiments with persons who were born blind that the
assertions of Goldstein and Gelb are contradicted by facts.
6
a In any
case, it is assumed here that the patient lacks all optical representa-
tions and that for localization he is reduced purely to tactile data.
Other experimental researchers working in psychology also reject
the question of tactile space.B4
We mention these individual scientific results because they are
significant for the present problem. If the question concerning an
independent tactile space is raised, then it can only be approached
through an abstractive separation of vision from touch. The
behavior of the blind provides an appropriate phenomenal basis for
the investigation. The anomaly, particularly of those born blind, is
such that they have already bracketed vision in a quasi-"natural"
way; in normal behavior such bracketing is a specific methodologi-
cal arrangement.
Nevertheless, the rejection of the tactile space from the pathological
side does not univocally mitigate against the existence of a tactile
space. In the evaluation of pathological discoveries, one is not al ways
consciously clear that the observer confronting the patient finds him-
self facing a radically alien consciousness. He can only describe the
63. See S. Monat-Grundland, who repeated the experiments of Gelb and
Goldstein on fourteen persons born blind. The work is published only as a
fragment and leaves the basic question open.
64. According to J. Wittmann, the blind person who does not move feels
"completely space-free" and "in no manner related to space"; he experi-
ences himself "as it were only as a thought function" (p. 433). Even his own
movements are not given for him as a change of place. Wittmann's student
W. Ahlmann reaches the result, on the basis of self-analysis, that the blind
person constructs his surrounding world purely temporally on the basis of
touch. A. van Senden, who gathered more than sixty reports from persons
with cataract operations, thought he could claim that the temporal schemata
of the blind have no spatiality. The spatial concept of the patients is to be
attained exclusively intellectually and without any sensory foundation.
Modally Distinct Sensory Spaces 129
patient's behavior with the aid of a privation, with the assumption of
a "breakdown" of normal functions; as a point of departure for de-
scriptive pathological work, this private approach is debatable in
contemporary research.
65
Again, it must be kept in mind that what is
presented here in "pure" description is always and ineradicably an
already interpreted comportment, especially when such a description
unavoidably in vol ves the observer's specific conception of space. The
complex of unavowed presuppositions leads to unclarity concerning
what one here simply rejects as "tactile space." lt is remarkable that
the researchers cited never explicitly raise the question of what is here
to be genuinely understood by tactile space. Obviously, abnormal
behavior permits two entirely distinct aspects of observation not to
be confused with the other. On the one hand, the patient is seen by
the doctor as located in the space of intuition pre-given to the doctor
himself. The doctor asks how the patient finds his way in it despite
a breakdown of the visual function, and how the space of intuition
is presented in the latter's tactile space. If one asks about "tactile
space" in terms of this aspect, then it is given as nothing other than
the space of intuition filled with tactile material. The nonexistence
of such a "tactile space" could then only mean a lack, or rather, a
failure of a spatial structural order determined by visual space, which
is alien in kind to the "pure" tactile manifold.
From this we can conjure up the affirmation as well as the
negation of an independent tactile space. There is a difference as to
whether one raises the question concerning the structural nature of
a manifold of positions merely accessible to touch in an otherwise
pre-given space (this question is extremely important for the psy-
chology of the blind; after all, it is of utmost importance for their
orientation), or whether one approaches this problem from an
entirely different aspect: the possibility of a spatiality constituted
primarily in tactility itself". This would have to correspond to a
manifold of places recognizable purely for its own sake as tactile
space, furnished with its own structural laws and existing without
the participation of the visual sense; in fact, it would not require the
65. Contemporary psychiatry-especially where it has assumed holistic
psychological aspects-wants to replace the concept of the "breakdown" of
the function with that of the "change" of the function, i.e., it regards the first
not only as a pathalogical datum but also as a basis of motivation for the
consolidation of the remaining functions into a new gestalt of activity.
Obviously this new conception of the "phenomenon" is on a new basis and
demands its critical revision.
130 Lived Space
visual sense at all. Indeed, in its own right it would be able to cast
sorne "light" on the specific structure of visual space.
The last point obviously leads to the genuine problem of a tactile
space. Its phenomenological treatment is hindered by the pre-given
interlacing of the tactile and the visual thing in all perception. At
first glance and from an appropriate viewpoint, it seems that it is
possible to establish the independent and essential features of touch;
it is possible to conceive of a touching corporeity that does not
require the activities of additional sensory functions for its opera-
tions. Yet such a corporeity is conceivable only for, and in terms of
a being who comprises a functional unity of tactility and vision. The
attempt to extricate the uniqueness of touch reductively from the
total structural activity of space-constituting sensory functions is
bound to the restriction that it can only be undertaken in terms of a
being who is simultaneously visual and tactile.
In touch the world is given in the primordial sense as standing in
opposition. Touch touches something, and indeed something mate-
rial, something thing-like, which it encounters and which simulta-
neously confronts it. This leads to the experience of the thing's
resistance. The tactile experience is not constituted by something
that is encountered "there" and "yonder," but rather by the direct
presence of something resistant. First, resistance is experienced in a
specific orientation: the object resists "at a specific place," be it
"there" on sorne objectivity orbe it "here" on one's own body-or on
both at the same time in the characteristic double givenness of touch.
In contrast, other experienced content requires a correlation with a
specific directing of attention. But the primary and exclusive con-
cern here is with the tactile manifold of "theres." Indeed, this
manifold is never detached from the lived body, although this
corporeal contact can be excluded from consideration and touch can
be investigated exclusively in terms of any tactile-spatial arder.
Something touched offers itself to touch primarily as a resistance
without intervals; there is no place on it that is not in principie
accessible to touch. This does not preclude that the touchable
something can have conditions under which touching is impossi-
ble--perhaps too high a temperature or an electric charge. This
conditionally determined untouchability is merely factual, given in
the specific empirical case, without constituting the limit in princi-
pie of the tactile activity. The suggested absence of lacunae in the
touched thing is present in touch even with optically discontinuous
structures (wire, lattice, etc.), since what here constitutes the "lacu-
nae" even in touch is precisely the continuously touchable connect-
Modally Distinct Sensory Spaces
131
edness of the edge.
66
This is not to argue that the thing must
necessarily be touched in continuity in arder to be present as
continuous. Thus the touching of a surface with spread fingers
reveals an experience not only of a multitude of discrete, singular
tactile positions, but also of the surface as a continuous surface. The
continuity of the tactile material-we use this term provisionally to
exclude tactile formal structures without wishing to lend matter
complete independence-is the first structural characteristic of
something tactile.
Nonetheless, the continuity examplified above is precisely what
can be contested. Obviously any co-experience of the total surface
in stigmatic touching can be traced to the in te grating role of a spatial
representation having an optical origin.
67
Certainly in normal
comportment the visual sense is of great significance. With the
exclusion of vision from the full complement of functioning senses,
then roughly speaking, the touching body can touch a surface only
stigmatically, and yet can have the experience of its connectedness;
the careful touch of the blind also seems to suggest that one can
speak of the connectedness of tactile material only through the
incorporation of vision. Yet the last observation speaks in truth for
an independent tactile continuum-the blind person does not touch
point for point because he doubts the continuity, but because he
wishes to inform himself about the size, form, and constitution of an
object whose continuity is already presupposed. But the other
observation, namely that even persons born blind are not
continuously engaged in touch for their orientation, does not speak
unconditionally for a tactile continuum. Pathological observations
are of little help in this question; clinically speaking, it is still
controversia! whether the disturbed are or are not in possession of a
kind of representational space. Nevertheless, it can be shown along
other paths that a tactile manifold of locations is constantly a
continuum-indeed, that the tactile continuum presents the
originary form of any possible kind of spatial continuity whatso-
66. From the above discussion it is furthermore understandable that the
results of radiology showing the discontinuous structure of matter lie at
another level of discourse. It has nothing to do with the continuity of the
phenomenal objects.
67. Thus argues, for example, H. Lassen (1), pp. 58 ff. For him all tactile
impressions are "embedded" in a "continuum of representational space."
Yet it seems to us that apart from Lassen's fundamental analysis, this
problem is not treated sufficiently far as a basis for rejecting a tactile space.
132 Lived Space
ever. Touch achieves this by virtue of its specific relationship to the
touching corporeity.
This cannot be sufficiently understood as a correspondence of
each tactile impression with a tactile sensation which, as a whole,
are linked together in the unity of the lived body. What is decisive
here is that the continuity of the tactile material is not given
simultaneously, as with visual material, but successively and,
indeed, in the succession of bodily movement. In this case the unity
of sensory activity and bodily movement is different and more
intimate than in the remaining sensory-functions. What is valid for
the ocularmotoric in vision is correspondingly valid for the
movement of bodily members in touch, yet this further has an
entirely different and more important meaning for the immediate
function of touch itself. In a completely originary sense, the tactile
sense is a motile sense. Something tactile is present as touchable
only and exclusively in actual corporeal movement. The arder of
tactile positions as tactile is exclusively an arder of succession. It is
given as such only in relation to the successive continuity of the
movement of corporeal members experienced from "within" not
only as a continuous series of tactile sensations, but also kin-
aesthetically as a continuous succession of corporeal phases of
movement. Here the externa! succession is simultaneously tied to a
successive inner continuity. But this "inwardness," as only one
particular mode of experience of one's own corporeal movement, is
always a continuous series of individual phases, resulting in the
property of continuity of that which is touched.
Thus it is apparently incorrect to derive the continuity of the
tactile material from that of the visual material. Rather, it is obvious
that the continuity of movement constituted in touch also founds the
continuity of the visual material.
The thesis rejected cannot be defended with allusions borrowed
from stimulus-psychology. This psychology claims that the dis-
tances between the points of sensation on the skin present "in
reality" a discontinuous material, and that continuity can be given
only with the support of visual function. Apart from the fact that
such argumentation moves at another methodologicallevel, it leads,
in its own way, only to a displacement of the problem. The
upholders of this thesis would have to draw the same conclusions
for vision and explain how seeing a continuous surface can result
from the mosaic-like structure of the retina, which in addition is
Modally Distinct Sensory Spaces 133
disrupted by a blind spot.
68
In all of these cases the physiological
facts do not lead to the understanding of the plain and simple
phenomena. The lived body, as a functional system, has its own
structure and arder of activity, which deviates from that of the
physical organism and is not comprehensible in terms of physiol-
ogy.
The unique interconnection between movement and touch leads
to further consequences and problems. lt must be added that the
continuity of the tactile material guarantees the continuity of the
total corporeal surrounding field. This means that the touching
corporeity has no location in its surroundings that is not basically
touchable or that could not be filled with tactile material. Each of its
constellations of limbs and each of its phases of movement can in
principie encounter something tactile; each inner kinaesthetic sen-
sation can be brought to coincidence with an external tactile
impression. Only one thing is essentially excluded: the surroundings
cannot be completely filled with tactile material, since this would
mean the destruction of the motile body.
What can be experienced of something touchable are not only
specific material properties such as hardness, softness, and rough-
ness, but also form and size. The distinction between primary and
secondary qualities is originally accessible to touch, and only the
former are relevant for the question of a tactile space. The touching
of size contains a first primitive determination of measure. Due to
the immediacy of corporeal contact with things, the quantity does
remain attached to a qualitative property; the form a.nd size of things
are here not only "related" to corporeity in its own orders of
magnitude, but are completely incorporated in the multifarious
qualitatively and intensively distinguishable, and distinctly felt,
movements of corporeal members. Yet a new difficulty appears.
Kinaesthesis is a multi-dimensional manifold, and as a sensation of
movement, it does not have a determnate direction. After all,
conditions and states are characterized by a lack of orientation
68. Even in the cases of hemianopsy, as is obvious from the researches of
W. Fuchs, there is no discontinuous visual field, but rather, despite an
increased scotoma, there are merely disruptions and displacements. Even a
partial hemianopsy does not yield an empty half-space but a newly
structured visual space, although viewed physiologically it should be half
absent.
134 Lived Space
altogether. How is it then possible to discover something like size,
form, or dimensionality of the tactile object by the sensation of
movement?
Indeed, our analysis is confronted by an insurmountable problem.
The frequently defended view that the manifold of kinaesthetic
sensations is reducible to three distinctive capacities connecting the
tactile and visual data borrows its argument from a region that is
aliento touch; this implies that there are not three measurements for
the purely tactile thing.
In addition, all touching is a flowing temporal event-which,
however, is in each temporal phase merely a touch of a surface. This
factual circumstance only seems to have an analogue in visual space.
Indeed, while there the thing is constantly seen only in specific side
surfaces and is first perceived from all sides in the temporal
movement either of the thing or of oneself, nonetheless each indi-
vidual aspect contains an indication toward further aspects. The
anticipation follows a strictly predelineated style. In contrast, the
individual tactile experience lacks similarly co-given indications. Of
course, successive touching takes place on an object in accordance
with specific expectations of further touch, yet here the resolution of
such expectation lacks the character of a mere fulfillment of some-
thing already anticipated. Rather, it is like an answer t a question
where initially the response is completely uncertain. Thus it is
necessary-and not only for the cautious touching in the fore-"sight"
of the blind-to follow up all anticipations with the actual process of
touch, surface for surface, form for form, in arder to obtain the total
object of touch-and this is in complete contrast to the object of
intuition.
Moreover, the purely tactile manifold of a surface structure has no
nearness or remoteness. In arder to appreciate this, one must
attempt to liberate oneself completely from optical conceptions.
Nearness and remoteness are not tactile factors since the touched
object lacks any spatial distance to corporeity. Each touched
objective there is at the same time corporeally felt touch here. Here
and there are thus not differentiated as places; rather, they are
distinctions in the directedness of experience. At the same time
there is no proper tactile perspective. There are no phenomenal
truncations and disruptions such as there are in visual space. The
tactile object can be arbitrarily turned without revealing tactile
adumbrations. This is confirmed by our previous explications of
perspectiva! appearances as objective phenomena in two senses:
they appear on the object itself as well as with another corporeity.
Modally Distinct Sensory Spaces
135
Hence they can only be given to a subject who in his corporeal
distance "objectively" has a world of pure oppositionality.
Perspective demands a radical separation of lived body and thing, of
corporeal conditions and objective reality; this separation is
unattainable in touch.
Moreover, perspective is related to spatial depth in the space of
intuition. Yet it would be premature to infer the lack of spatial
depth from the absence of perspectivity of the tactile manifold.
Obviously here the question concerning tactile space attains its
proper importance. Indeed, touch lacks the phenomenon of things
being covered over and replaced by other things; the tactile thing is
where it is and is constantly accessible to touch. Taken strictly
phenomenally, there is no such thing as one touchable thing being
in the way of another. Yet it is remarkable that the sketches of those
born blind do in fact show coverings and intersections, even if there
is a complete lack of perspectiva! presentation. Thus there appears
an impression of depth in the plane of the sketch. It is already
striking that these patients master the problem of sketching and that
for them it makes sense to present the touched world on a surface at
all. Such mastery assumes the differentiation of their tactile world
into surface depth. Yet contrary to first impressions, these
discoveries do not lead us far into the question posed. Since such
sketches are capable of various interpretations, we are of the
opinion that they can neither prove nor contest an originary "tactile
depth."
69
To decide this question, we should investigate a circumstance that
did not receive due consideration in our previous deliberations: the
specific manner in which one's own corporeity is experienced in
touch. Closer observation of the resistance by virtue of which
"something" is first present at all to touch reveals its dual nature:
experience not only shows that the tactile thing resists corporeity,
but that the touching corporeity resists the touched thing. Corporeity
69. It is to be noted that the researchers who have dealt extensively with
the sketches of those born blind have specifically rejected tactile depth. W.
Voss assumes an a priori spatial representation of those born blind as an
explanation. H. Lassen (1) attributes to the blinda spatial field from a visual-
like structure, despite the exclusion of the eyes, since according to him
vision is nota peripheral process in a sense organ but a central event. In any
case, for him the pictorial depth in the sketches of the patients remains
exclusively visual; the "picturability" of the world can never be compre-
hensible from the structure of the tactile world.
136 Lived Space
thereby experiences itself with respect to its touchability: it is a
"physical body," a thing arnong things-and yet at the sarne time, it
rernains unbridgeably separated frorn thern in that it is a feeling body
and, indeed, one that feels "frorn within. "
70
This phenomenal datum
is quite important. Such inner awareness of one's own body reveals
at the same time an arder of depth. Depending on the kind of
impression and its location on the body, tactile impressions on one's
own body are experienced more or less peripherally. The intensity
and quality of the effect are not only given as such, but are traced at
various depth levels. This experience reveals something crucial,
something that cannot be mediated by optical experience, and as a
pure state or condition cannot be accessible to vision; it is founded
exclusively in touch. The touching of one's own body grounds the
originary distinction between surface (as touched "outer" surface)
and depth (as a specific experiential datum of one's own feeling-felt
corporeity), and thus a first conception of the "body" of an "object"
in its full voluminosity. Insofar as the touching body is also a
touched body, "objects" are structured at the outset in accordance
with surface and depth. lt is not surprising that this depth of tactile
space differs from that of visual space. lt is not perceived in the
objective arder of things, but rather is experienced primarily in one's
own corporeity. lt is a sensory function that in its objectivating
activity remains halfway between outer and inner. lts objectivity
remains embedded in the medium of changing corporeal conditions.
The experience of depth in tactile space, deviating from that in
visual space, explains, on the one hand, the lack of the phenomena
given in visual depth (adumbrations of form, coverings, etc.)7
1
and,
on the other, the positive fact that a lived body functioning and
acting exclusively in terms of touch can have a space structured in
terms of surface and depth-and at the same time, know how to
represent it on a pictorial surface without the additional assump-
tions of a purely intellectual spatial intuition, of visual residua, etc.,
becoming necessary.
The above analysis does not claim to have dealt exhaustively with
the problem of visual and tactile space. Their characteristic
70. Concerning the constitution of the body and the experiences of outer
and inner in so-called double sensations, as well as with respect to the mode
of givenness of one's own body, see pp. 143 ff. of this work.
71. All perceptual phenomena are also lacking in the inner givenness of
one's own body.
Modally Distinct Sensory Spaces 137
properties were sketched only to the extent that they were essential
for the problem of the founding of the space of intuition. Visual and
tactile spaces, separable from the space of intuition only
abstractively, are accessible to descriptive analysis only condition-
ally. They ultimately lead to methodological difficulties befare
which a phenomenological procedure in the sense previously used
comes to a halt.
SECTION TWO
Questions of Space Constitution
Chapter One
Corporeity and Spatiality
1. Methodological Survey
The preceding investigations remained within the framework of a
phenomenological method. This procedure makes the claim that it
can acquire its data from the reflective orientation toward various
modes of comportment of the corporeal subject in space. What was
present for description was primarily not space asan object given in
thetic space-consciousness of the subject, an object that would
provide the basis for the judgement that space is; rather, what was
described was space as it becomes appropriated in pre-reflective
corporeal comportment-in execution. This point of departure was
based on the claim that this relationship of the subject to space is
more originary than, and is assumed by, the oojective relationship.
This assertion is not without difficulties. While we believed it
necessary to maintain a linear progression from one spatial form to
another, we were also clear that although this movement was
phenomenologically justifiable, it was at the same time valid only as
a foreground. In truth, the correct way of thinking dictated by the
state of affairs should resemble a spiral, insofar as at any given stage
of his being spatial, the subject must have already traversed all
spatial modes. But this subject is no one else but ourselves. In this
regard, the phenomenologically descriptive work unavoidably had
to be disrupted by an occasional self-reflective move.
138
Corporeity and Spatiality 139
Furthermore, we were aware that what was phenomenally "given"
-if at first glance it was most obvious-was not something that carne
to us of its own accord and could be received by us without raising
the question of its origin. Indeed, the sense of the concept of
givenness was alloted its full due. We took something as present for
phenomenological explication and description in contrast to its
being constructively devised or deductively derived. Nevertheless,
at the very outset this given was seen from a specific aspect that
already determined its choice. The comportment of the corporeal
subject in space remained the leading point of orientation. In attuned
experience, in goal-oriented activity, and in sensory intuition, the
subject appeared in three different modes of orientation toward the
world, with their respective sense-orders, each of which turned out
to be constitutive for the structure of its own space.
It was stressed at the outset that the anticipated division and
classification had to proceed abstractively, in a very specific sense,
for the sake of clarity. The analysis had to separate and to thematize
what in reality was present as a single moment in a concrete totality
of factual behavior. The attuned, the active, and finally the sensibly
intuiting lived body, cannot be extricated purely in themselves.
Rather, the lived body is to be regarded as lived body insofar as it
functions in the ways we have described, and even then never purely
and exclusively in any one aspect. Correspondingly, it is also true
that the three "spaces" analyzed are not parts of a single space but
merely structures of the one space according to the corresponding
modes of corporeity. Furthermore, the work of analyzing found itself
in the circumstance that we were concerned with an unseparated but
nevertheless clearly articulated unity of distinguishable structures
and this became the methodological justification of our procedure.
There are additional reasons why this investigation is called
phenomenological with certain reservations. In a specific sense
every phenomenological endeavor is precursory and points to an-
other endeavor that cannot remain with the phenomena as such.
Even the concept of phenomenon involves a metaphysical position.
That something is "appearing" makes it depend on something
whose appearing it is; it is appearance in relationship to being. It is
irrelevant for us whether the latter, for example in ontological
realism, "announces" itself in the phenomena, or whether it is
"concealed" by the phenomena and thus justifies hermeneutical
ontology. All attempted separation between a phenomenology on
this side of metaphysical positions and the metaphysical positions
as such could not be maintained. While dealing with the subject
140 Lived Space
matter, each method progresses factually. Yet what is immanent in
the essence of any originary method is the meaning of being in light
of which the subject matter is explicated.
For the following investigation this means that the descriptive
analysis of space necessarily includes an ontological treatment of the
problem of space. The ultimate question is not how space is
experienced, but what it is. This is the reason why our investigation
had to hint at this question, and to hint at an answer ahead of time.
Finally, in our attempt to provide an answer we consciously accept
all reservations and restrictions resulting necessarily from the meth-
odological situation suggested above.
2. The Lived Body and the Physical Body in their Relationship
to Space
In all the controversia! questions concerning the relationship be-
tween the subject and space, it appears certain that space is somehow
"related" to corporeity. In the previous analyses this type of rela-
tionship turned out to be twofold: on the one hand, space appeared
as the "wherein" of corporeity and thus as a condition of its possi-
bility for motion; on the other hand, space appeared to be primarily
conditioned by, and structurally dependent on, the manner and mode
of corporeal movement. This correlative relationship of corporeity
and space, seen up until now in the three forms corresponding to the
modes of comportment of the corporeal subject, must subsequently
be illuminated more closely with respect to its own structure. The
various modes of spatial appropriation present in the different types
of corporeal modes of comportment must be traced back to the con-
ditions of the possibility for a constitution of space as such.
This kind of regressive procedure reveals an unavoidable point of
departure reflected in two concepts whose meaning is guaranteed
and established only in their mutual relationship. Corporeity and
spatiality are as abstractions totally empty and are never given as
data. They reciprocally require one another for their meaning. We
cannot ask how one emerges from the other or how one is founded
by the other, but only about specific correlations within their
reciproca! relationship itself. One does not precede the other-
neither ontically-genetically nor ontlogically. Corporeity was not
"before" spatiality, since to be corporeal, i.e., the foundation, is to
assume spatiality; the latter, in its own turn, was not "before"
corporeity, since corporeity determines spatiality as what it is-
spatiality of corporeal movement.
Corporeity and Spatiality
141
That corporeity is already in space is a conception of natural
consciousness having sufficient justification in the simple phenom-
enal state of affairs. Corporeity of a determinate nature and mode of
functioning, and corporeity of a determinate concretion as "lived
body" is found in a spatiality that is likewise of a concrete determi-
nate structure with an already "finished constitution." Thus
corporeity finds itself in a space that for natural consciousness
appears as given.
But can we think of the lived body as being-in in analogy to the
way one thing is contained in another, and understand such being
contained from a specific mode of corporeal functioning? Can it be
surmised that in the nature of its functioning the lived body is
posited as the ground for the possibility of apprehending itself as
being "in" another?
Let us recall the conception of the lived body as it was sketched at
the outset. As it is present in nonscientific understanding, the lived
body presents itself as a completely indifferent psychophysical unit.
Consequently, reflective analysis cannot see in it anything more than
a complex unity of activities and functions. This unity is the
determining basis of its being no other except this entity with its
modalities, i.e., with its comportment toward the world. This unity
turned out to be tripartite: corporeity appeared in its sensible
comportment in the modes of attuned experience, goal-oriented
activity, and sensory intuition.
However, these modes of comportment do not reveal the lived
body as primarily present to itself; rather, what is given to it in the
first place and above all is a section of the world. Where the lived
body itself becomes given, it does so in a way comparable to any
other mode of givenness. In sensation as "corporeal feeling" it is
present in a specific, qualitatively and intensively unique state. This
state does not signify objectivity. It is an awareness (Innewerden) of
the lived body in a nonthetic mode of consciousness which, how-
ever, has two specific characteristics. First, corporeal consciousness
is consciousness of the lived body as a whole. Even with strictly
localizable organ sensations, the organ is not sensed alone; rather,
the lived body as a whole functions as the phenomenal background
of all these single sensations. Secondly, each actual consciousness of
one's own lived body already belongs to the latter's entire existence
(Da-sein) prior to all differentiated states. Each corporeal conscious-
ness always and already assumes the prior being of the entire lived
body which is not exhausted by the singular given states. If feeling
(Befinden) is the way in which one becomes conscious of one's own
142 Lived Space
lived body in the mode of having, as a mode of being of the subject
it is nonetheless already prior to any consciousness of "having." The
lived body is a founding and basic phenomenon for any kind of
"having" as such; as will be shown more precisely, it is the first and
necessary condition for the possibility of any form of possession of
something or of knowledge of something.
The certitude of the lived body, mediated by its feeling, is a typical
certitude "from within." But where is the "outside" of this "inside"?
How is it that we make this spatial distinction between inner and
outer? The boundary between inner and outer runs visibly at the
surface of the lived body. Yet this expression is senseless-a
functional unity cannot possess a surface. Lived body, in its strictly
phenomenal sense, is on the hither side of all separations between
inner and outer, as it is hither of all separation between physical
and psychic. That there is such a distinction is founded in it. It is
corporeity itself, then, that constitutes this separation, and indeed
through its specific function of touch.
It is now necessary to investigate tactility not with regard to tactile
space, but with respect to the mode of giveness of one's own lived
body. In touch there occurs the originary communication of the
corporeal subject with a world "outside." Something resistant is
encountered in it, and the lived body is confronted by something
material and thingly. But how can a thing confront the lived body?
Obviously the thing does not resist it, since it is present to the
touching corporeity without limitations. Resistance cannot mean a
limitation of corporeal activity; moreover, in the experience of
resistance the tactile function reveals a determination characteristic
of the lived body, namely its being a physical body. The lived body
as a material thing, as a "physical body," constitutes itself through
a specific corporeal function. Here the founding role of the lived
body for any kind of givenness becomes clear. Any theory that thinks
that it can construct the lived body on the model of levels-material
and the psychic (Husserl), or the inorganic, the organic, and the
psychic (N.Hartmann), whereby the physical body ought to be the
founding level of the remaining "higher" levels-fails to conform to
the phenomena given. Even when they claim to be phenomenolog-
ical, they uncritically follow the conceptions of the special sciences.
The essence of originary corporeity, being nothing else than a system
of functions, does not include being a merely "physical" body. It is
possible to think of a corporeal being whose functions are so designed
that a physical bodily world, and thus corporeity as a physical body,
would never cometo appearance. At the same time, the corporeal
Corporeity and Spatiality
143
being would not cease to be corporeal. To be sure, such a corporeal
being is not our own. My lived body is physical corporeity; never-
theless, 1 know of the existence of a physical thing called m y "body"
on the basis of my lived body. As a physical body, the lived body is
given in immediate connection with the material world. Yet the
material world was not befare the lived body nor did the latter
precede the former. The constitution of a resistant thing-manifold
and the constitution of the lived body as a physical body are only
two sides of one and the same process.
But corporeity, as a physical-bodily corporeity, is not yet suffi-
ciently determined. If it is material, thingly like other things, then it
must not only be touching but also touchable corporeity. lt is
characteristic of the tactile sense that it has unique sensory function
of the so-called "double sensations." The lived body is a subject as
well as an object of touch. In its resistance it is given to me
objectively, like other things, and at the same time it is experienced
subjectively as receptive of sensations. The possibility of this imme-
diate union of coincidence between corporeal feeling and corporeal
perception is the ground for the differentiation between outer and
inner. Yet this union of coincidence must become present as such to
awareness. This requires a specific experience of identity wherein
corporeity, in its touching knows that it is identical with itself as
touched. 1t is only to the extent that this consciousness of identity is
there, that the concepts of outer and inner assume a fulfillable sense.
1t is not that 1 somehow touch a thing, but that it is my corporeal
thing-which while touching is given as self-touching-that makes
this originary separation between inner and outer possible. With it
the lived body has constituted a first spatial relationship, which is
retained in all higher functional activities. But insofar as the lived
body establishes this activity only in a mutual interplay with the
world of things, it itself becomes exposed to them. On the basis of its
physical-bodily constitution the lived body has all the attributes of
things: extension, form and size, materiality and weight. All thingly
qualities are at the same time corporeal, in the precise ontological
sense that one can speak meaningfully of them only with respect to
the strict correlative relationship of lived body and world. Thus
there are no corporeal bodies or bodily things, no size or form of
one's own lived body, without the possibility of its presence in the
world of extended things and vice versa. At the same time, however,
there is a place in space for corporeity justas there is for things. The
being-in-space of the lived body solely as a spatial thing is, to say the
least, a delusion of a naive consciousness; an essential determination
144 Lived Space
of the lived body to the extent that it is a bodily corporeity, is to
constitute a space-establishing function as such.
But while it is comparable to things, this does not mean that the
lived body thereby becomes a thing among things. A thing does not
experience its thinghood in its encounter with others; it does not
have any "thing-experience" of "itself." Yet the self-experience of
the corporeal "thing" is entirely different in kind from experiencing
all other things. While grasping myself as a body, 1 am present to
myself "in the flesh" in an incomparable way. Thus the sensory
intuition of other things are given in continuous syntheses that fuse
the partial intuitions of perspectiva} aspects into an intuition of a
whole-this is essentially excluded for my body, since my body can
never become fully accessible in the successive progression of a
perceptual series. But there is no need of such syntheses or identi-
fications, such fulfillment of indications and anticipations of the
particular parts. Rather, my body as a whole is present to me
immediately and actually; it is not given objectively but felt through
an arder of depth clearly differentiated in accordance with quality
and intensity. And this felt state is first constituted in the touch of
one's own touching-touched lived body.
lt is important to see that touch, in its space-constituting activity,
is not repeatable by any other sensory function, not even by vision.
The latter is neither constitutive for the materiality of the thing nor
for the differentiation between inner and outer. A non-touching
corporeity would not have an external world. Indeed, it seems like
tedious speculation to say anything about the world of a lived body
that is essentially not a touching lived body. Nevertheless, sorne
hypothetical statements are possible. We must exclude both the
mode of corporeal givenness in vision and the constitutive role of
touch for vision that was already demonstrated.
No corporeal feeling, no inwardness corresponds to vision alone.
There are no visual sensations, as there are sensations of pressure,
hardness, and warmth. Where such sensations occur in vision, for
example with a dazzling and harsh light, vision becomes disturbed.
One's own lived body is not given in any manner in normal vision;
1 do not see "in" the eye in the same way 1 touch "at" the skin; rather,
1 see into space. (This mode of expression already presupposes the
accomplished distinction between inner and outer.) The eye is an
organ of my corporeal body insofar as it is receptive to touch and
pressure, but not insofar as it sees. As seeing, the eye is a member of
the lived body, but it is not itself given to me in seeing. Only the
oculomotoric kinaesthesia, as a sensation of an organ, gives me a
Corporeity and Spatiality 145
vague and merely peripheral inner awareness of my corporeallived
body. Yet there is an absence of the unity of coincidence between
one's own seeing and seen corporeity, between subject and object;
this is possible only in touch. Even kinaesthesis already assumes the
lived body as touchable, and thus assumes the tactile function.n
Indeed, the lived body is visually distinguishable, but in a way
that is different from touch. As visible, m y lived body is a lived body
for others who perceive it, but nevera lived body forme. This is not
only because it is only partially perceivable by me-even in this
partial perception it is aliento me in a unique way. It is conceivable
that this partial view can be acquired by the use of a mirror, whereas
there is in contrast no analogously functioning implement for touch.
The mirror which "lets me see" my lived body, in truth does not let
me see it, but only its image-and this only through the mediation of
an already pre-given external world. Mediated in this manner, this
image can deceive, distort, or flatter, can lead to doubts whether it is
really my lived body whose image I perceive. In contrast, touch is
beyond such doubts. A touch, a blow, immediately convinces me of
the mineness of the lived body. Yet it convinces only me. Were the
other to touch my corporeal body, he could have the same tactile
impression as from other things, but never "my" sensations. In a
strict phenomenological sense, there is no co-feeling of corporeal
sensations.
The functional uniqueness of touch appears still with yet another
characteristic. What it encounters is given to the lived body only
successively; the touched thing first unfolds in a succession of
individual tactile impressions. But what is the ground of the
experience of the unity of the touchable thing in the temporally
discrete succession of tactile experiences?
The very sense of the question implies that the touched thing is
already known as one. This knowledge is attained not merely
through one's own movements but through self-movement or, more
precisely, through the motion of a self who knows itself as moving.
To grasp something as one thing in the flow of impressions is to
assume that this flow is apprehended as a movement "in" which
something identical is given, because it is appreheiided as a move-
ment of something identical: the moving lived body. Thus this
72. In fact, it is difficult to say something about a merely seeing corporeity,
not even about that which it sees. Any perception of something extended
would already be impossible for it; seeing would not even be seeing of
surfaces. Vision without touc,;h remains completely incomprehensible.
146 Lived Space
question is addressed at the outset to a corporeity that is not only
active in touching and is identical in touch, but that in touching also
knows itself as the same and, at the same time, knows itself as
self-moving.
73
One must be cognizant of the unacknowledged "presupposition"
that the genuine sign of life is self-movement. The unreflective
consciousness refuses to grasp movement otherwise than as a
relative change of place of something identical "in" time. It can
hardly conceive that a movement or rather, that what it grasps as
such could also be apprehended as a temporal change "in" space;
although this is thinkable with difficulty, yet that it be sois rejected
by unreflective consciousness even when the change indicated
shows no sign of preference for either orientation. Nevertheless,
when the process is grasped as a movement, as a change of place in
space, it is because the change is accepted on the assumption of the
identity of something in motion; it is a movement for a
consciousness that grasps identity. This is the meaning of the
frequently stated proposition that all motion "requires" an identical
carrier of the movement-the process, which could be grasped as
movement, "requires" identity; the latter presupposes the identity
of consciousness with itself, a consciousness of a self maintaining
itself in the changes of the perception of movement. Only such a
73. This is not to claim that there is a reflection inhering in each
experience of self-movement such that this identity is given thetically to
awareness. lts coming to attentive perceiving does not have the character of
being-conscious-of-something, as is appropriate for every reflectio:q; "di-
rect" self-consciousness is not at all an intentional consciousness. The
structure of the experience of movement is not as if in it the moving
corporeity were meant and in its differing phases of movement were
established in its identity. Indeed, a reflection of a specific kind can intend
the identity of an ego-corporeity, yet this expressed intention is first possible
on the basis of an unreflective pre-knowledge of one's own identity as it is
primarily experienced in motility. Perhaps it would be more appropriate to
speak here of ipseity and not of identity. This term is appropriate for the
unreflective selfhood of the ego as an identity given here. Identity presup-
poses difference, and strictly speaking, is never a simple experience, but
always a result of acts of thought that are not expressed; since they are
oriented primarily toward differences, they establish sameness despite the
difference, or rather they question and test the sameness of something that
is being asserted as identical. The ipseity meant here, however, is not
questionable or testable with respect to identity, but is rather a simple
phenomenal datum.
Corporeity and Spatiality 147
consciousness can experience change and transformation, as well as
their differences.
The notion of the self-identity of the lived body is contained in its
motile comporting itself toward the world even if no knowledge of
this self is present. Animal corporeity may be present to human
understanding as self-less and, thus, privative in contrast to a self.
However, human corporeity not only essentially determines itself as
the lived body of a self, but also knows itself as such; it determines
itself as an ego-corporeity and thus it is primarily a self-moving
corporeity in the genuine sense.
It was already suggested that movement as such could be indiffer-
ent to space-time (p. 41). This cannot and does not mean that it is
"outside of" space and time; it is grasped only in both determina-
tions. Yet in both, movement indicates not only the duality of both
determinations but also their originary unity. This is still present in
tactile activity. With the constitution of outer and inner that founds
a primordial spatial relationship, the touching lived body attains the
contiguity of the tactile there only in a succession of now and now;
it knows how to attain a manifold of locations,, a space in a
succession of individual sensations. While bound to the identity of
a feeling-felt self, they are grasped primarily as a succession, as a
forward movement in time. By constituting space in the mode of its
own tactile movement, the lived body also constitutes time in the
manner of ek-stasis; yet because it finds itself in this kind of
self-movement, this space-time unity is already doomed to disrup-
tion. While the lived body apperceives its originary mode of func-
tioning as self-movement, simultaneously it lends the movement the
sense of a change of location of its body (which remains identical
with it), i.e., of its temporal process in space. An assumption of a
continuous transformation of its body in time would be tantamount
to the dissolution of its self. Space-time is ontologically relative to a
corporeal being as such; space and time are ontologically relative to
a lived body that is also an ego-corporeity.
Without scrutinizing the specific function of time more closely, it
is essential to trace the relationship between space and corporeal
movement at greater length. We attempted to decipher this relation-
ship in its ontological origin. We saw how the lived body, with its
inner and outer aspects, founded a primordial spatial relationship
and at the same time revealed its own fundamental spatial determi-
nation as corporeal body. The incipient aporia of corporeal being-in
can be resolved only in this way: while as a lived body it constitutes
space, through a specific corporeal function it constitutes itself si-
148 Lived Space
multaneously as a physical body in space. Thus corporeal space is
that which is determined by its activity. The paradoxical structure of
this state of affairs is founded on the fact that here we are confronted
with the ultimate relationship beyond which thought cannot
progress.
If touch turns out to be the incomparable function that grounds the
strictly correlative and reciproca! relationship between body and
space, nevertheless it does not solely comprise the richness of this
relationship with regard to its contents. While founding every
sensory performance, it remains at the same time half immersed in
this performance. Due to the immediate contact of the lived body
with things through touch, the lived body cannot accomplish
objectification in tactility alone. By presenting an externality alen to
corporeity, it founds an initial objectivity by setting it against its own
subjective inwardness; yet corporeity cannot accomplish this sepa-
ration from itself since in touch it is present to itself in a unique way.
The existence of space, which in its fullness of structural character-
istics transcends the mere tactile manifold, requires further func-
tions and forms of activity of corporeity. Indeed, the space of
intuition has its sensory foundation in touch and the space of action
also has a constitutive moment in it. Yet in their own arder and
articulation they do not exhaust themselves in being merely a
modified tactile space. The thing as ready-to-hand, as well as the
thing present-at-hand, is not comprehensible solely as a set of
sensible constitutive moments. Corporeity in a situation is ego-
corporeity,lived body of a self, and thus its space is not only a space
of corporeal functions but also a space of its performances of
consciousness.
The preceding investigation is left with a specific difficulty:
whether and how is it possible to unify the fact that, on the one hand,
space is a space of movement "for" as well as "through" corporeal
movement, and that, on the other hand, it is an object of conscious-
ness. Thereby the problem already arase asto whether the solution
to the problem of space could be found solely in the objective form
of intuition, whether the specific being of space could be found in
the categorial activity of the subject.
3. The Lived Body and Consciousness
The meaning of the problem raised above could be formulated in
an alternative way, i.e., whether the corporeal subject is originarily
a lived body or whether it is originarily consciousness; whether, in
Corporeity and Spatiality 149
an ontological sense, the lived body precedes consciousness or
whether consciousness must rather be a requirement for the being
not simply of corporeity but of ego-corporeity.
The obvious epistemological problem that emerges here must be
bracketed. That consciousness has a gnoseological priority requires
no discussion. That the lived body, in order to be known as lived
body, assumes consciousness, is an analytical proposition. There
can be no controversy concerning the irreducibility of what pertains
to consciousness itself. In order to derive consciousness from
something else, consciousness is required; any attempt to reduce it
to something else involves a petitio principii. Consciousness as such
is irreducible; it is something that has itself for foundation and must
be accepted as such.
This does not seem to be the case with the lived body. If it is the
lived body of a consciousness, then the latter is that from which it is
what it is. We have done justice to this notion above, insofar as we
have never regarded the lived body as something simply self-
sufficient and to which consciousness was merely "also" attributed;
from the outset we regarded it as the lived body of a subject, and
showed that even its most elementary functions, such as locomotion,
present corporeal being as a consciousenss.
While its ontological relevance was admitted for the being of the
subject, there was nevertheless no prejudgement concerning the
possibility of an ontological derivation of the subject's corporeity
from consciousness. Even though at the outset the so-called
philosophy of consciousness is designed to be essentially epistemo-
logical, it apprehends corporeal functions as dependent in their
being on the grace of something else, and does so in two ways. Thus
in his Ideas and in the Cartesian Meditations oriented exclusively
to "pure" consciousness, Husserl not only dissolves the entire
region of extra-corporeal being into an intentional being in
consciousness, but also submits his own lived body to the decree of
the transcendental-phenomenological epoche. The lived body is,
then, only something constituted in pure transcendental conscious-
ness while the latter-in its manifestly unlimited and uncondi-
tioned intentionality-remains the sale absolute. Thus one can
arrive at such statements as these, that the subject "localizes
himself" through his lived body, and that there is a possibility that
my lived body does not exist. Even Heidegger, obviously bound to
the transcendental-idealistic heritage, speaks of Dasein's spatializa-
tion in its "bodily nature." This view especially comes to the fore
150 Lived Space
whenever a specific problem is excluded from the domain of
interrogation. 74
Pursuing the problem of spatialization, Merleau-Ponty is forced
into an opposite conception. In his Phenomenology of Perception,
the lived body attains an ontologically fundamental position. While
inappropriate as a principie for deriving consciousness, it is elevated
to the status of an unconditional being. Although this seems to offer
a decisive and clear view, Merleau-Ponty's presentation moves in
formulations favoring the other extreme of the ontological relation-
ship between lived body and consciousness: "Il y a done autre sujet
au-dessous de soi, pour qui un monde existe avant que je sois l et
qui y marquait ma place. Cet espirit captif ou nature1, c'est mon
corps ... (There is, therefore, another subject beneath me for whom
a world exists befare I am here and who marks out my place in it.
This captivity of natural spirit is my body ... ").
75
This conditioned
relationship, as it is conceived here, makes the short-circuited
thought comprehensible. The latter consists of Merleau-Ponty's
belief that he can completely explicate the structures of the
corporeally-centered space of the "etre engag" and make this space
comprehensible to the corporeal subject without somehow introduc-
ing objective space into discussion.
By attempting to avoid such truncations of the problem, we also
recognize a specific methodological difficulty in our own investiga-
Han. It appears that our problem lies in the particular difficulty of
appropriately conceiving the being of the subject both as corporea1
subject and as corporeal subject.
Consciousness does not localize "itself" as a lived body in space--
insofar as it is consciousness it was always consciousness of a lived
body. It was never without a lived body. Dasein does not spatialize
itself in its corporeity as though, prior to and over above it, it had
performed differently-it was always spatial in the form of being
corporeal. At the same time, lived body as my lived body, and hence
experienced as mine, was not "befare" or "beneath" but was always
mine, was always the lived body of an ego. Without me it would
never have begun, and in accordance with its meaning it was always
ego-corporeity, the lived body of my self-consciousness. Even with
strict adherence to the ontological significance of such temporally
sounding expressions, the relationship that prevails here is such that
74. E. Husserl, Ideen Il, 22 ff.; M. Heidegger, p. 108.
75. M. Merleau-Ponty (1), p. 294.
Corporeity and Spatiality
151
as soon as it is expressed it becomes tantamount to the disruption of
the true state of affairs. As the paradigm of a strictly reciproca!
implication, it resists any effort to grasp its fundamental structure
linguistically.
This not only means that consciousness is not the sale irreduc-
ibility, justas the lived body is not the sale contingency, but rather
that this twofold non-derivability of the corporeal subject is at the
same time bound in the ontological unity that consists of a recipro-
cally conditioned relationship. While this may be obvious with
regard to the lived body, the reverse, the determination of conscious-
ness through corporeity, will be opposed whenever gnoseological
considerations cometo the foreground. As a relationship of being,
the one-sided and irreversible epistemological orientation from
consciousness to the lived body is irrevocably two-sided. Even
where pure consciousness recognizes corporeity as corporeity, it
does so only through a consciousness already containing a moment
of corporeity itself. The only "pure" consciousness is one that
functions independently from all merely factual, individual condi-
tions of a lived body, but not from the facticity of corporeity as such.
A consciousness without corporeity is not only not discoverable, but
not even thinkable. Consciousness would be a completely empty
concept because its total activity cannot be determined without the
inclusion of corporeal functions. Even its complete composition
with respect to its contents-the noetic as well as the noematic-
contains corporeal implications. This is not only valid for a non-
intentional consciousness whose really intrinsic contents are, as
simple sensations, not merely mediated through the lived body, but
are the very contents of the corporeal states and are themselves
corporeal. It is also valid for objective consciousness, whose inten-
tionality comprises only a specific accomplishment of subjectivity as
such. That all factual consciousness is intentional-and that in
particular it is doxic and thetic, i.e., in its objectifying acts it consists
of various modes of conceiving, positing, interrogating "Being"-
does make present, in the total complex of the performances of
consciousness, a specific contingency that resists any further inter-
rogation and is not derivable from the subject's corporeity. Yet even
here we find no acting being whose really intrinsic as well as
intentional contents did not have certain foundations in the prior
givenness of corporeal-sensible factors-and these factors do not
reveal indices toward sorne ultimate acts that would have to be
deciphered in arder that we may comprehend the nature of such
factors.
Chapter Two
The Space of Movement and Objective
S pace
1. Spatial Structure and Corporeal Facticity
The separation of the lived body and the physical body and the
insight into their relationship has solved the problem of "being-in"
in the following way: corporeity as a bodily corporeity in space and
corporeity as first constituting space, should not be seen as two
exclusive states of affairs. Thus all that remains is to answer the
following question: how is space, as a space of movement, related to
an objectively conceived space? Obviously this question is not
identical with the one dealt with befare. It is not concerned with the
problem of the correlation between corporeity and space (which can
be explicated from an anthropomorphic viewpoint for the animal
kingdom), but rather with the specific relationship of space as a
form of fulfillment of corporeal movement, on the one hand, and on
the other, space as space of objective consciousness.
The former is co-determined by the latter. Despite its methodolog-
ical difficulties, this insight could not be avoided even by phenom-
enological investigations if they intend to offer an exhaustive anal-
ysis of the states of affairs. The space of movement was not
comprehensible without consciousness of objective space. The latter
turned out to be underivative. This led to the notion that fundamen-
tally it is the only one relevant for the ontological question of space.
If it is the sale space illuminating other "spaces," then ontological
significance ought to be attributed only to it.
This would be the case even if the being of the lived body were
dependent on the grace of consciousness and if the latter were the
sale unconditioned aspect of the subject. Yet corresponding to their
mutual non-derivability, it is impossible to reduce the space of
movement to the space of objective consciousness and vice versa. At
the same time, corresponding to the mutual implication of lived
152
The Space of Movement and Objective Space 153
body and consciousness there is a reciproca} relationship between
the space of movement and objective space. What constitutes the
structure of the latter, under closer illumination of the conditions of
its possibility, turns out to be a sum of determinate properties. As
already emphasized, the latter are not deducible from the corporeity
of the subject without going through the reflexive circle. Yet objec-
tive space is conceivable only as co-determined by a corporeal being.
In modern and contemporary philosophy, there are many attempts
to investigate rationally the ontological status of space, to demon-
strate its structural components and to deduce it from various
premises. It is striking that these arguments reach their results at a
price of a complete glossing over of the historical dimension.
Homogeneity, void, and boundlessness appeared in these arguments
as "apriori" determination of space. Yet in historical retrospect these
concepts have assumed predominance only since the beginning of
the Renaissance and only became accepted as the universal con-
sciousness of space in the seventeenth century. In view of this fact,
our claim about "the" space acquires a specific significance. Al-
though we have not defined space in any strict sense, we suggested
that it has one of the fundamental conditions in the corporeity of the
subject. And in view of the historical transformations of space-
consciousness, it is necessary not only to differentiate the prominant
and enduring problems, but also to test the historically given stock of
phenomena.
If one were to observe the multitude of typical arguments concern-
ing the question of three-dimensionality with respect to their com-
mon denominator, then it would become obvious that despite the
different points of departure, all are traceable back to factors that are
based in the ineradicable conception of the subject's being as a
corporeal subject.
If one wanted to derive three-dimensionality "logically" from the
essential connection between surface and depth, as is done by C.
Stumpf, orto grasp it with H. Lassen in the unity of two "funda-
mental moments" of extensionality and intentionality, one would
discover that the process of proof ultimately finds its starting point
in the facticity of the corporeal modality of being.7
6
Lassen
76. Sturnpf (1), pp. 85, 275, explicitly rejects rnovernent asan "integrating
condition of spatial representation," although sornewhat later he adrnits that
the representation of depth "is least extricable frorn rnovernent." While
dealing purely psychologically with the conception of space and at the sarne
time, confusing it with Kant's pure forrn of intuition, Sturnpf has in view
154 Lived Space
nevertheless contests "that the question concerning the degree of
dimensionality could be answered in terms of the existence of
distinctive spatial directions." "Space as a pure manifold must
always already be given befare the question of its number of
dimensions could acquire sense." Las sen correctly states that strictly
speaking the number of dimensions cannot mean the number of
distinctive directions, but rather the degree of manifoldness. Since
we are at the pre-quantitative level, this characterization has an
advantage in our investigations. It can avoid metric characterizations
(the highest number of perpendicularly related lines, etc.) and can
deal with the topological properties. lt is instructive how Lassen
determines the degree of manifoldness of space by its precise
derivation from the "intuitive-topological" conception of space.
That spatial depth constitutes a degree is shown by the univocity of
its direction, allowing only one possibility of uniting points located
one behind the other.
77
Since the corresponding operation on the
surface offers various possibilities, the degree of surface manifold-
ness must be at least two. That there can only be two results from the
fact that the manifold of surface relationships of a region is so
constituted that given a sketch of straight lines on the surface, there
can be at most two straight lines that could connect just once in such
a way that none of the remaining straight lines are intersected.
7
s
Thus Lassen's three-dimensionality is proven. For him it is not "an
accidental property ... but rather a formal, dosel y structured system
of relationshi ps," . . . an "a priori fact."
Here, however, its necessity can only offer the possibility of an
insight into its fundamental conditions from which the topological
states of affairs mentioned can be understood. The latter are obvi-
ously based in the corporeity of the intuiting subject. The proof of
three-dimensionality can make sense only to a sensibly intuiting and
only the moving thing; yet even this presupposes the self-movement of the
perceiving subject. Concerning H. Lassen see (1), p. 35. We subsequently
support this presentation since it appears to us to be most careful and,
within the framework of our statement of the problem most, instructive.
(The proof for three-dimensionality has been established by Lassen for the
perspectiva! space of intuition. Yet in terms of its numerical dimensionality
it cannot deviate from the homogeneous space, since it is the manner and
mode in which the homogeneous space is attained in its monadological
aspect.)
77. The univocity of depth will be discussed subsequently in the context
of the problematic of the visual ray. See pp. 248 of the present work.
78. lntuitable topological examples are given by W. Lietzmann.
The Space of Movement and Objective Space 155
thus self-moving being. Who else could understand that spatial
surface and depth "necessarily" demand one another? Nevertheless,
the converse is also valid: only the lived body as corporeity of an
intentional consciousness can perceive the sensibly intuited things
in the arder of surface and depth, extension and intention. In
addition, it seems that an interrelationship between the degree of
dimensionality and the directional oppositions is unavoidable.
Three-dimensionality, as well as its orthogonal relationship, can be
founded only on the elementary orientations. Indeed, it was already
pointed out that in the space of action the three-dimensional,
rectilinear system of dimensions is not yet given with the elementary
pairs of opposites. While it is impossible to treat the elementary
directions of orientation in accordance with a system of coordinates
of "the" space given in any other way, the former, in turn, are are not
yet such a system. Nevertheless they contain the founding moment
for any spatial coordinate system as such.
In any case, an attempt to deduce the triadicity of space from
kinaesthesis misses its aim. Kinaesthesia is purely a qualitative and
multi-dimensional manifold. If one were to attempt such a reduction
by recourse to certain coincidences and couplings of sensations of
movements, resulting in an exact arder characterized by triadicity,
one would be moving in a circulus in demonstrando: in truth one
posits a merely hypothetical connection that cannot be shown imy-
where, while the space of movement as triadic manifold has been
already articulated in sorne other manner. That it can and must be
seen as a three-dimensional arder does not depend on an uncondi-
tioned "pure" space-consciousness. The latter is, and continuously
remains, a consciousness of a lived body in space and is conditioned
by the lived body. But that it is precisely this lived body, with these
movements, that the lived body is constituted in this and no other
manner, possessing these factual members and functions, is irrele-
vant for the triadicity of space and the orthogonality of its dimensions.
W e saw that the determination of the elementary differentiations
of orientation was only possible through the inclusion of corporeal
organization on the one hand and through the functional activity of
its members on the other.
The univocity of this determination could be maintained only
with the aid of the principie of economy with which the active
corporeity complied in its orientation toward the world of the
ready-to-hand-the most originary form and gestalt of its intention-
ality. The same holds true for the very attribution of all six orienta-
tions of movement to three originary and permanent pairs of
156 Lived Space
contrary oppositions: right and left, above and below, front and rear.
They are originary in that they are given on the basis of anticipated
activity, but not primarily as "pure" spatial orientations belonging to
sorne kind of theoretical intuition. Their correlation is apprehended
precisely as fixed and interchangable, and in this correlation the
correlated have the character of opposing one another. The equal-
ization of these opposites, the homogenization of the directional
differentiations, takes place in self-movement which relativizes the
oppositions. Instead of the three pairs of opposites, there appeared
only three functionally equivalent (constantly directed to the fore)
frontal orientations in the space of movement, three "dimensions" of
frontal movement in which the original oppositions are dissolved.
The system of three spatial dimensions is completely conditioned by
the corporeal moment and the supposedly "pure" consciousness can
loca te the triadicity of its space only because this space is at the same
time the space of a functionally determined corporeity organized in
this and no other way.
A consciousness without corporeity would have no knowledge of
the "perpendicularity" of the dimensions. The orthogonal relation-
ship is indeed not topological but metric. With this we can only deal
subsequently, specifically since such a relationship does not belong
to the composition of the phenomenon in the sense previously
delimited. Yet it will be shown that this manner of measuring angles
must be understood as a formalization founded in the oppositional
directions of movement.
Particularly difficult problems seem to result in our investigation
with regard to the relationship between the space of movement and
the space of objects, especially when one considers the boundless-
ness of the latter. We are deliberately avoiding the concept of infinity
in its proper sense of "open" infinity. Subsequently it will presenta
specific difficulty; meanwhile, we continue to speak of the infinite
in an imprecise way as the and so on of space. Natural consciousness
can only grasp it in this manner. What is decisive at this point is
obviously that this "and so on" of space is not comprehensible either
through the corporeal structure of the subject or through any other
datum. After all, the insight of the previous discussions showed that
the objectively grasped space cannot be regarded as a sum composed
of singular spaces of motion; rather it is always presupposed in
phenomenological observation. This phenomenal state of affairs
must be confronted with our conception of the reciprocally condi-
tioned relationships between the space of movement and the space
of objects.
The Space of Movement and Objective Space 157
Phenomenologically speaking, we face a difficulty. The modality
of givenness of "the" space for natural consciousness is debatable
prior to all theory. Whether "the" space is the finite but limitless
space of antiquity or the open and infinite space of the modern
consciousness cannot be decided, since as a whole it is not given
adequately in any intuition of a single particular. lts limitlessness,
whether conceived in one or the other sense, can in no manner be
brought befare objective space-consciousness. This contrasts with
the finite, singular space of intuition in the perspectivally intuiting
consciousness; its givenness "moves" concurrently with the clues of
the "continuation" of space. This conception is supported by
kinematic structures and is exhausted at the achievement of an aim.
A more refined characterization can be readily found in the
Aristotelian conceptual schema of possibility and reality, of poten-
tia! and actual infinity. Although originally Aristotle brought this
schema to the treatment of the Zenonian aporia of continuity with
regard to the "inner" limitlessness of divisibility of a finite stretch of
space, the conceptions are also applicable to the problem of spatial
limitlessness toward the outside. But what does the frequently
expressed claim mean that space is "potentially" infinite"?
If one accepts the Aristotelian dynamei on as a concept opposite to
the energeia on or entelecheia on, then the possible could be taken
to mean the not yet real; but the possible can only be conceived
insofar as it can be converted into reality. Thus understood, the
validity of possibility cannot be excluded and its realization is its
positive characteristic, enabling a conceptual differentiation of the
possible from the impossible.
Yet for dynamis in the sense meant here, these conditions are not
fulfilled. After all, the infinity of space is not a type of possibility
that could ever become a reality. The potentiality of space is a
"pure" possibility that could ever become a reality. This state of
affairs might have been what Scheler meant by space as a pure
capacity-in his view a capacity for self-movement. At the same time
he regarded it as non-real and attributed its purported reality to the
"deception" of natural consciousness. That this latter claim contains
difficulties is shown solely by the fact that a deception, if it should
exist, must be able to be demonstrated, and that in principie it can be
abolished in the context of further or other experiential or epistemic
interrelationships. Since this is excluded in Scheler's case, it must
be a "deception" of a specific kind; more precisely, it must be a
mode of experience that cannot be grasped with the concept of
deception.
158 Lved Space
Yet Scheler's point of departure can be evaluated positively if the
concept of dynamis is taken in a second Aristotelian sense. It need
not be taken as opposite to energeia, and it need not mean here the
possibility of teleological realization; rather, it may simply mean the
"capacity" to do something or to passively undergo something.79
This is a characterization of potentia that when applied to a specific
experience of the capacity for movement, can offer useful solution to
our question.
Since all factual corporeal movement constitutes only a limited
spatial stretch, it takes place on the background of consciousness of
"further" possibilities of movement. In a concrete phenomenological
sense this possibility affirms a progression of continuous movement
of corporeity that is free in principie and simultaneously the
"continuation" of space. Yet through every factual fulfillment, this
possibility of experience is not varied, diminished, weakened, and
finally extinguished; rather, insofar as movement continues at all,
this possibility remains intensively as a constant presence. This
means that the possibility of progression "into infinity," precisely as
this pure possibility, contains in itself the moment of infinity. As a
possibility "toward" infinity it is itself phenomenally infinite and
never disruptible possibility. As an attribute of space, infinity is in
truth nothing other than a characterization of my possibility of
movement.
This does not yet necessarily determine the open-endless exten-
sion of space. What is characterized here as the infinity of intensive
movement is fundamentally open to two kinds of structures of
"infinity" of space-the boundless but closed, turning back upon
itself ("not genuinely infinite") and the open-endless ("genuinely
infinite") space.
Both structures are fundamentally compatible with the specific
character of movement's infinity. With regard to the genuine infinity
of space, it is necessary to have a specific mode of experience of the
type of movement which, despite its periodicity and rhythmicality
(as merely an endless repetition of the individual steps of move-
ment), is experienced as movement forward in the full sense of the
word. It is only with and in the experience of moving forward that
the background of further possibilities of movement is differenti-
ated -in accordance with the "already" realized and the "yet" to be
79. An exhaustive treatment of the Aristotelian dynamis cannot be offered
within the framework of our investigation. Compare the initial exposition of
this problem in W. Wieland, pp. 292 ff.
The Space of Movement and Objective Space 159
realized possibilities, in accordance with what has gone by and what
is to come-it is only by "flowing onward" that time, along with
space, can be infinite in a genuine sense.
This does not mean that the great historical turning from the topos
of antiquity to the empty, infinite space of modern times could be
understood solely in terms of corporeal relationships of movement.
What has determined the conception of space concretely throughout
history has never been a phenomenological presentation of corporeal
modalities of movement; rather, they were conceptions based on
world views and metaphysics. Indeed, until Newton's intuition of
space they were thoughts laden with theological content. But this
does not contradict the claim of our exposition with its clearly
outlined limits. Our intent is to present the fundamental conditions
of possible space-constitution as such. Such conditions must neces-
sarily be fulfilled if space is to be given, although their. mere
fulfillment does not offer sufficient guarantee for a specifically
determined topological spatial structure.
Befare discussing the question of structure, we should make sorne
additional remarks concerning the character of infinity. That my
possibility of movement is itself infinite suggests that the infinity of
space cannot be an intended aim of my movement. This state of
affairs may have been the source of the notion that infinity is
"merely" potential. Thus potentiality appears to be a lower ontolog-
icallevel than actuality; however, this misunderstanding can arise
only if the meaning of potentiality is understood solely in analogy to
actuality. Yet we are dealing with a possibility that is of a lesser
mode of being not because it can never become reality, but precisely
because in its uniqueness it has an ontological function that cannot
be confronted by any modal counter-concept.
Furthermore, the potential infinity of space does not mean that
space is "possibly" infinite. The experience of limitless possibility
of movement is in its own way not merely possible; rather, as
experience of potentiality, it is actual, i.e., it is an experience that is
actualizable in each factual phase of movement. Space "is" indeed
infinite only in the mode of potential infinity, although potentiality
in turn is not something merely possible. It is not a contradiction if,
on the one hand, potential infinity is attributed to a space of limitless
possibility of movement and if, on the other, such an attribution
appears as an actuality of awareness. The potentiality will not
thereby become actuality-this would be counter to it-although
space possesses in its characteristic of potential infinity a unique
actuality in the objective consciousness of space. This complicated
160 Lived Space
state of affairs may be the catalyst for the succession of constantly
rejuvenated theories engaged in the controversy about the actual or
potential infinity of space. At the same time, it calls for a specific
type of insight that does not bring space to consciousness in the same
way as objects or things, but rather allows space to remain in its
phenomenal uniqueness, the mode of givenness which is inaccess-
ible to closer description.
2. The Problem of Empty Space
The previous investigations have shown that space is where
corporeity is. They have also shown that as self-moving, corporeity
constitutes space, and moreover it can first have space conceptually
and objectively at all through consciousness of this capacity for
movement.
In any case, the consciousness of the capacity for movement is not
identical with the consciousness of space; in the everyday
experience of space, this space is already conceived as pre-given
and as independent of one's own movements. But the specific
experience of space by natural consciousness has its own
foundation in the manner and mode of possible movement. While
the corporeal subject intuits itself as motile, it experiences at the
same time that space is there in the movement, that each actual
movement has space at its disposa1. As has been shown, however,
this experience of movement is so ordered that each factual
movement is directly and always known from the background of
broader-more specifically, previously realized and yet to be
actualized-possibilities of movement as a continuum of movement
phases. Within this continuum, the movement just accomplished is
only one phase-one singled out as phenomenally given in the
mode of full presence.
In this context the experience of an already executed movement is
significant. The essence of continuity of any movement includes not
only the capacity of optimal "forward" progression, but also the
experience that any chosen movement can be repeated "backward."
This means, to say the least, that an absolute beginning of movement
can never be given. In any case, there is a beginning of having
moved, and there is a first instant of completed movement, but there
is no first instant of being in movement. Each beginning of a
movement is itself already the continuation of movement that in turn
must have its beginning, and so forth. The experience of beginning of
movement would destroy the experience of movement itself; phe-
The Space of Movement and Objective Space 161
nomenally speaking, movement is only graspable when it is already
movement.
But if all movement is "already" movement, then space must
"already" be if movement is to take place at all. This allows us to
understand why the corporeal subject in his natural consciousness
or attitude conceives of the space of his movement as a pre-given
existent. For him, space is something that takes place in, but not
something that first becomes through movement,ao This conception
is as little a delusion as his continuity of movement would be. That
each factual movement already occurs in space corresponds experi-
entially to the fact that movement has always and already consti-
tuted space. Of course, we are in no position to grasp the beginning
of this constituting activity. Indeed, the mere possibility of such an
original determination would abolish movement itself and destroy
space.
"In" space does not mean that space must necessarily be con-
ceived in accordance with the view claiming that the relationship
between the moving body and space is one of an encompassing
emptiness containing a body. Space understood substantially as
emptiness, existing independently of-although containing-
things, is a relatively recent conception. lt should be emphasized
that this characterization is not at all the datum of pre-reflective
everyday consciousness of space. If one adheres strictly to what is
given, then there is nothing that could readily be designated as a
receptaculum rerum; rather, the latter is an aspect of a specifically
modern theory of space (whose historical genesis will not be
discussed here any further). Empty space does not exist in the
thought of antiquity. For Aristotle, space was not conceivable
without a body. The concept of an empty space was incompatible
with his physics. lt is remarkable that Aristotle establishes his theory
of topos not on the basis of speculative deliberations and intellectual
constructions, but in relationship to his theory of movement. What is
methodologically remarkable is that in his theory of space, Aristotle
touches on the direct experience of everyday life more extensively
than all subsequent thinkers. However, what is thematic in Aristotle
is only the place (topos) of a body as its limit (peras). Place is a region
of a specific nature, with its own power of attraction, toward which
each body must tend in its natural movement. By way of natural
movement, each body has an appropriate place in accordance with
80. This holds analogously for the experience of "still" in the future
being able to move oneself for which there will be space.
162 Lived Space
its essence ( eidos). Thus for Aristotle the determination of place
includes the determination of essence. Indeed, the topos here is
completely distinguishable from its various occupants. It is nothing
like a thing, and in contrast to things, it has its own meaning-but
only to the extent that the place has a dynamic influence on the body
and the influence is not identical with the influenced, not to the
extent that the place can exist independently and indifferently to the
moved body.
81
81. According to M. Jammer (p. 17), Physics Delta 4 210b 34-211 a6 should
demonstrate "the reality of space." This interpretation is anachronistic or at
least misleading in its mode of expression. When topos is discussed, and it
is said that topos is periechon echeino on topos esti, chai methen ton
pragmatos, all that is said is that topos is nota property of a body. Yet it
would be a mistake to seek in Aristotle an intimation of the modern concept
of space. Indeed, the place is separate from the body (choriston), but this
detachability ought not to be understood in terms of the later conception of
an empty, indifferent space in contrast to things; rather, as is stated in
Physics Delta 1 208 b4-11, it must be conceived as a distinction between
thing and a place ( . . . tonto on ton eggegnomenon chai metaballonton
eteron panton einai dochei ... ). This is to be interpreted in terms of the
Aristotelian theory of movement; the "self-sufficiency" of the topos in
contrast to the body that finds itself in it consists of the influence of the
topos on the thing: deJonsin oti esti ti o topos all oti chai echei tina dynamis.
Recently, in her interpretation of Aristotelian theory, H. Conrad-Martius
has pointed out certain ambiguities and contradictions (pp. 109 ff.). Yet
these, as well as the critical claims that Aristotle did not have a sufficient
concept of space, seem to us to be groundless. lt would be inappropriate to
accuse Aristotle of simply having a "false conception of physics," justas it
is nonhistorical to miss a specific chapter in his Physics concerning "space"
(as translation of chora), yet at the same time to accept that in the
Aristotelian theory of topos "space is always meant there in a secondary
way." In accordance with an Aristotelian way of working, the meaning
would have been explicitly mentioned and analyzed if the term charos had
had the specific meaning that is being attributed to it. The casual and
non-terminological usage of this word *e.g., Delta 1 208 b7 and 32 suggests
that it did not mean something like space "in contrast to" place in the sense
of modern terminology. lt is remarkable that in this sense the Greeks did not
have an expression for space. That Aristotle speaks of topos and charos but
not of space, that he does not have this concept of space, characterizes his
"space"-conception. Similar misconceptions are found in J. Cohn, who
accuses Aristotle of "lacking a useful concept of space" (p. 41) and traces
this lack of understanding to the Zenonian aporia, claiming that Aristotle
"has not made space sufficiently clear" (p.42). Indeed, Cohn thinks he can
The Space of Movement and Objective Space 163
This opposition between the space-cosmos of antiquity and the
empty space of the modern physical sciences constitutes by itself an
essential point for any theory that wants to assure for the latter space
the dignity of an apriori necessity, of a timeless essential arder of
pure consciousness of something similar. Here to raise and to decide
the quaestio iuris in favor of one or the other conception would be a
senseless undertaking. The only legitimate attempt is to illimunate
how it is possible to think the substantivization of an empty space,
which-as already mentioned-does not even become thematic for
natural consciousness. Yet its thematization in science may appear
to be motivated by the natural consciousness of space. These motives
may once again be traced back to the originary space-constituting
strata of corporeal subjectivity.
It would be beneficia! to recall the originary mode of activity of
corporeity that founds the primordial spatial relationship through
the constitution of inner and outer. It turns out that touch is not only
responsible for this relationship, but that it is also constitutive for
any possible sense of emptiness.
Touch reveals the unique mode of functioning of corporeity insofar
as this is essentially related to its movement. It is not as though this
capacity does not "belong" to other sensory functions; yet here it
appears to be most closely connected to a function of performance,
since it can create a manifold of theres only in actual self-movement.
What is more important is the reversal of this state of affairs: even if
not given thetically to consciousness, each singular phase of motion
of the corporeal body constantly includes a touched there and there.
The reciproca! relationship between the tactile manifold of places
and self -movement illuminates the continuity of this manifold; while
this manifold is constituted through self-movement, it is presented
as nothing other than the continuity of this self-movement. This
means that there is no place in the surrounding field of the touching
corporeity that is not basically tactile, i.e., that could not contain
tactile matter. At the same time, every phase of its movement and
constellation of its limbs has in principie the possibility of encoun-
tering something tactile. Each kinaesthetic sensation from within can
come into coincidence with an external tactile impression. The con-
tinuity of movement allows for the fact that the manifold of theres
constituted through it is never discrete.
show that Aristotle has committed a "basic error" of "confusing space with
body" (p. 114).
164 Lived Space
Only one thing is essentially excluded: the manifold cannot be
seamlessly filled with the tactile "object." Such filling would be
equivalent to the destruction of the lived body and of the tactile
space itself. That this continuous filling never takes place is not a
metaphysical accident; rather, it is the touching lived body itself that
in its movement first grounds the manifold of theres. The continuity
of the tactile space does not mean that it is solidly filled in; it means
rather the capacity in principie to be filled in the succession of
constituted tactile manifolds. Thus tactile space yields the phenom-
enon of emptiness, more precisely that of the empty place. Tactile
emptiness does not merely presenta self-sufficient datum insofar as
it is not a visual or in sorne other way determined suppositum;
rather, it is also that which founds the significance of any concept of
spatial emptiness.
82
Insofar as the lived body becomes a physical-
bodily corporeity and thereby founds the primordial spatial relation-
ship, the lived body at the same time discovers itself in space. This
relationship necessarily contains emptiness as the "space" for the
possibility of corporeal movement. In its ontological origin, space
thus determines itself essentially as empty space.
The emptiness founded in touch is nevertheless only "in-
between-space." It is not an absolute emptiness preceding objects,
but is relational with respect to things. In a continuous change of
place by things, each empty place can be grasped as a possible
location for an object and conversely, each location may be regarded
as a possible empty place. Yet it may mean that this is all a matter of
"mere" conception which can arbitrarily focus on the one or the
other aspect. Or, to speak with Aristotle, it may appear as if the
matter is left to a conceptual choice as to whether we see space as a
totality of places or whether this totality is primarily obtained from
empty places "for" possible bodies. These two conceptions are not
equivalent. They become equivalent only when in the first instance
the Aristotelian topos is divested of its specific place character, and
when the second case explicitly assumes the additional notion "for
possible bodies." The conception of the equivalence of all places
demands a fundamental transformation of antiquity's theory of
movement. This was done gradually in later physics. It was the
82. This is most clear in the visual field which lacks emptiness. Without
disruptions, color joins color, form joins form, and the boundary of one
figure is already a boundary of another without there being "space" between
any two of them. The latter emerges only where not the field but space is
grasped, i.e., in the movement of things.
The Space of Movement and Objective Space 165
Newtonian conception of gravitation lhat ultirnately offered a strict
formulation of a law. However, the strictness was "only" postulated,
without experientially revealing the uniform rnovement of gravita-
tion of a thing. The thing had to be conceived specifically as a
"physical body" in arder for the Newtonian conception to be an
adequate expression for the strict homogeneity of space.
The equivalence of all places for movement expressed here
simultaneously leads to the possible conception of its indifference to
the things. If location and empty space are no longer distinguished in
terms of rank, but are only distinguished factually through the
existing constellation of things, and exchangeable with its changes
and if, in addition, the distribution of things is constantly discrete
while their movement is continually cohesive, then their places
must appear to be no more than momentarily "occupied" places of
a continuum which, taken by itself, is regarded as completely empty.
Yet actually it is not this emptiness, conceived "in terrns of an
interconnection with the world of things and their movement," that
leads to genuine difficulties. What is problernatic is its absolutiza-
tion. This conception assumes an empty space preceding and
existing independently from all thinghood. While the role played by
this "absolute" space in modern physics was tried-even Newtonian
mechanics did not use it-it has obviously been influential, through
Kant, in determining the scientific consciousness of the nineteenth
century. Kant thought that if we were to think away everything from
space, the latter could not be thought away and hence exists for
itself. Indeed, Kant attempts to defeat the realistic conception of
absolute space through his Copernican turn-yet what is significant
is that even in his transcendental-idealistic conception of space as a
pure form of intuition, space is still conceived as an independent
emptiness preceding all things.a3
Methodologically speaking, it is remarkable in our context that
Kant starts with the conception that "one" could never imagine that
there is no space while at the same time "one" "can think that there
are no objects to be encountered in it.'' Obviously Kant is arguing here
on the basis of natural consciousness and develops his proof for the
apriority of space by criticizing this consciousness. Yet this point of
departure, attributed frequently to Kant, is by no means self-evident.
And if "one" would wish to claim that Kant meant natural everyday
extra-scientific consciousness-Kant himself does not use these con-
83. I. Kant, Vol. III, Critique of Pure Reason, "Transcendental Aesthetic,"
Section 1, p. 2.
166 Lived Space
cepts-then what is attributed to it by Kant as a content of represen-
tation could at least be contested phenomenologically. In any case, it
is clear that since then, this "container fiction" has been presented
as an inviolable possession of the natural consciousness of space-
At the same time it is also strange that phenomenology, educated in
the subtle analyses of the contents and modes of consciousness, sees
this "container-conception" asan indubitable phenomenological da-
tum of consciousness.
84
It seems that here phenomenology has
uncritically posited as phenomenal "givenness" what one, so to
speak with Kant, can only think. And this is possible when one, like
Kant, "thinks" of the Newtonian space of physics-but not when one
traces precisely what is phenomenologically manifest in extra-
scientific space-consciousness. Even where space is given thetically
for this consciousness, it will not be given asan empty space in the
sense outlined above. There is a clear distinction in meaning as to
whether things, moving freely in space, permit a conception of emp-
tiness on the basis of their free movement-i.e., its determination is
attained privatively from the world of things and becomes a spatial
residuum so that this emptiness assumes a basically relational char-
acter-or whether one attributes this emptiness as a positive deter-
mination of space, existing independently of all thinghood, and lends
ita substantial meaning. As already presented, the latter is a scientific
conception whose basic motive is found in the natural consciousnss
of space, but which is nota conception of this consciousness. The
container fiction, as a phenomenological datum of the natural space-
consciousness, is itself a fiction of the phenomenology of natural
space-consciousness.
This does not remove the primary difficulties. If space is con-
ceived as a container of things and hence itself seen as a thing, then
it must be in space, etc. Such a conception would lead toan infinite
regress. Yet this conception, attained through a process of iteration
and leading to absurdity, is nowhere to be discovered phenomeno-
logically. Whenever it is advanced, it has to do with a polemically
exaggerated "fiction" of a reflective critique attempting to replace
the "mere" unreflected natural consciousness of space with a "true"
conception of space-without, of course, explaining its appearance.
Thus it fails in its obligation.
But how, in the final analysis, is the relationship between a body
and space presented? What is truly found phenomenally in natural
space-consciousness?
84. E. Husserl, Ideen 1, Ideen 11; M. Scheler (3); N. Hartmann (3).
The Space of Movement and Objective Space 167
Primarily, this consciousness readily distinguishes the spatial
determinations of the thing from space itself. For it the thing is not
identical with where it is found. If this "wherein" were to be
regarded as a kind of a more encompassing thing, then there would
be no information given by the thing concerning what accommo-
dates it, apart from the sole judgement what the thing is "in" it. In
turn, however, the previous analyses as a whole were oriented
toward spatial things and their distinct modes of being and, correl-
atively, toward the specific modes of movement of corporeity
required to read the structure of space from them.
The sole fact that space is indeed graspable with things already
shows that even natural consciousness is concerned with the rela-
tionship between thing and space, a relationship that is fundamen-
tally distinct from one obtained between two things. The latter mode
of being-in is determined by a relationship of two or more things;
containment in-in accordance with its intuited fundamental mean-
ing-is an inter-thing relationship. Yet the latter must already be
spatial, must be in space; a containment of one thing in another
already presupposes space as a foundation. That which founds a
relationship cannot be one of the relata. Although graspable "on" the
basis of the given interrelationships among things and determinable
"from" it, space is not this interrelationship itself. We speak of
extension, form, size, and movement with respect to things, and
location, place, and region accrue to them. Yet space itself does not
possess these characteristics; it is neither extended in accordance
with position and place, nor does it move. Rather, its own structural
characteristics will be attributed to it, although they will not
coincide with those of the spatial.
In our discussions, the concept of representation was used in
various ways to designate the relationship between thing and space.
It seems that in contrast to the previously ennunciated characteriza-
tions of the relationship, and with its closer delimitation, this
concept is more appropriate to the phenomenon in question and
simultaneously it avoids sorne main difficulties. Generally speaking,
the relationship between the representing and the represented is
neither between something and that which encompasses it, nor
between a form and its content, both taken to be independent; rather,
it is a relationship unique in kind.
It implies that space, represented by a spatial thing in it, is not
given immediately, in "the flesh"; at the same time this lack of
presence "in person" does not mediate space as if it required sorne
kind of intermediary conceptions or indeed.modes of deduction in
168 Lived Space
arder to get to space from the side of things. Rather, space is "there"
immediately with the spatial things. It is presented through things
and given only as presented "through" them. It is apprehended and
must be apprehended solely in them. It is not given in any other
mode apart from being represented in spatial things. It has its being
essentially in being represented through its contents, justas they are
spatial only through it.
This avoids a substantial distinction between space and thing. The
inherence of things in space is thus different from that of contain-
ment of one thing in another. Things are not just in space, space is
also in things. Both have a meaning of inherence that must be
non-spatial in kind. Hence the mode of space as being in things
enables the reified spatial containment of one thing in another. This
corresponds with the assumed relationship of representation; the
relationship between that which represents itself and the repre-
sented is in fact not a spatial relationship.
In our opinion, only this and nothing else ought to correspond to
the phenomenal composition of natural space-consciousness. But
when one attempts to conceive of space as open-endless extension,
this infinite space is constantly a space ofthings conceived as linear,
unlimited and mobile, but not a space conceived "primarily" as
empty and "then" as filled with things. Even here space can be only
if it presents itself in things and thus is graspable in its structure only
from them. The state of affairs is thereby established that space
presents itself in things-and this "that" must be taken and themat-
ized as its own existence and mode of being. But it is one thing to
take space as space of things in simple infinite progression, and
another to take it as a substantial, infinite emptiness and to think it
independently of things. In the latter view, space appears as a result
of an abstraction whose ultimate foundation, as has been shown, can
indeed be found in the corporeal-sensory phenomena. This abstrac-
tion, as has been obvious in the historical development of Newton-
ian mechanics presupposes, nevertheless, a fundamental transfor-
mation in the conception of a thing. This transformation leads from
the direct perception of an "intuited thing" to a specifically be-
stowed meaning of "physical body," with all the constitutive deter-
minations alien to the simple thing of intuition (regularity of
relationships in terms of mechanical causality, reduction of qualita-
tive characteristics to quantitative, etc.).
One can quarrel about the physical usefulness of such a space and
reject its existential justification for physical theory on the very
grounds that this theory stems from it, thus demonstrating that it is
The Space of Movement and Objective Space 169
superfluous. But what is decisive in this context is that geometry, as
a science of space, first became possible with the conception of
empty space as such.
In its transition from the Egyptian art of calculation and measure-
ment, where the purely mathematical and intuitive are completely
interwoven with the quantities derived from things, to Greek math-
ematics and its method of proof, the history of mathematics offers an
impressive transformation to the strict mathematical thinking. Nev-
ertheless, in Greek science, geometry could not have been concerned
with space itself. The irreconcilability of the ancient conception of
space-which found its culmination in the Aristotelian theory of
dynamism-with the consequences of the geometry of Euclid-
leading to a purely mathematical, homogeneous, and infinite
space-is perhaps ultimately the reason why euclidean geometry
was not also conceived as a geometry of "euclidean space." Further
development was required befare geometry was not only concerned
with geometrical formations in space, but constituted the concept of
an independent mathematical space.a5
3. Concluding Observations on Lived Space
The discussion above has reached the limits of what belongs to the
problematics of the phenomenology and ontology of lived space.
Concerning the latter, let us offer brief conclusions with respect to
the most important results within the framework of our investiga-
tions.
Space is, insofar as there is corporeity. Originarily, the latter is
neither in space nor outside of it; rather, corporeity is spatial in the
mode of space-constitution. The being of space is relative to a
corporeal being: in its given structure, it points to correlatively given
formations and concretions of corporeal movement. But space is not
exhausted in this correlation. The correlation of corporeal function
and spatial structure is intertwined with that of corporeity and
consciousness. Beyond the corporeal correlations, the latter has a
characteristic of being an intentional consciousness. In accordance
with the intertwining of this twofold correlation, the space of
movement and objective space present two abstractly distinguish-
able, yet de facto mutually conditioning moments of space-
constitution. In its full concretion, space is ontologically relative to
85. An essential although brief summary of this development is offered by
J. O. Fleckenstein.
170 Lived shace
a corporeal subject whose being consists of the ambivalence of
corporeal being and the transcendence of corporeity by an inten-
tional consciousness. If space is not exhausted in its being related to
a lived body, it nevertheless exhausts itself in being a space for a
corporeal subject.
This relativity does not designate a mere epistemological relativ-
ity. That space as an object of knowledge is related to a subject of
knowledge is a tautological statement-that in its being it is related
to the being of a corporeal ;ubject and is limited by such a
relationship is a metaphysical statement of vast importance. It
includes a departure from ontological realism, which attributes to
space an existence absolutely independent of the subject, such that
the relationship to a subject would be completely additional, exter-
nal; space would have its being befare all consciousness, being what
it is whether there is consciousness or not.
The very conception of space could be understood only as a
depictive repetition in consciousness. With this is bypassed the
demonstration of an image-consciousness which must still be
searched out in a theory oriented phenomenologically to the lived
content of natural consciousness; this conception would in addition
have to explain how the space of intuition, seen merely as a
subjective distortion of the homogeneous space, is possible as the
space of senseful comportment and meaningful corporeal orienta-
tion. Moreover, a further problem to be solved by the theory
suggested is the following: the depicturing relationship of image and
object is transposed into the relationship between knowledge and
space, a relationship that is itself spatial and thus presupposes
spatiality.
That these difficulties can be circumvented by strict adherence to
the claims of natural consciousness and what it irrevocably pos-
sesses should have become clear from the preceding investigation. It
did not contest the claims of "natural realism" with respect to the
conception of space-it contested it so little that it sought to
legitimate such a realism by providing an insight into its motive and
origin. Yet a retrogressive movement to its transcendental conditions
leads in turn to consequences incompatible with ontological realism
as a philosophical position. To attribute a transcendental turn to an
idealistic metaphysics could hardly be justifiable exept under un-
conditional maintenance of one of the traditional alternative
schemas whose validity should be discussed with respect to the
point of departure of this investigation.
D ~ s p i t e the differentiation of their metaphysical explication of
The Space of Movement and Objective Space 171
being, idealism and realism have their starting point in the problem
of the externa! world that transcends consciousness. Yet if they are
to be understood as final products of phenomenologically-descrip-
tive efforts, represented most recently by the theories of Edmund
Husserl on the one hand and Nicolai Hartmann on the other, they
remain related less in the diametrical opposition of their explication
of phenomena than in the place of origin of their search for
phenomena. Both assume the (thetic) consciousness of being. In one
case, being as being-phenomenon is a constituted being in con-
sciousness while, in the other case, the being phenomenon is
precisely a phenomenon of being, i.e., is taken as being in itself. For
Husserl's constitutive idealism with its bold claim of "eo ipso being
the true ontology," it is valid to say that being's transcendence of
consciousness, as a mere transcendence of consciousness, is more
precisely an intentional transcendence "in" consciousness. Con-
versely, for Hartmann, who claims equal subtlety of analysis of
phenomena, all possible positions apart from the self-evidence of
realism are excluded at the outset-and, on the same grounds as
Husserl, Hartmann claims that the transcendence of being in con-
sciousness is a factual transcendence.B
6
We shall not trace here the basis that ultimately leads both to
one-sided interpretations of being and of being "in" consciousness,
and to prejudiced standpoints. Although initially they do not admit
their theoretical positions, the subsequent formulations of their meta-
physical conceptions appear more like mere explications of positions
held from the start. It is of little use to pursue here the question
whether the alterna ti ve between idealism and realism can be solved-
or whether each attempt to give a final and valid decision concerning
this question must founder on the aporia of intentionality.
We turn now to our problem: whether space has a being transcen-
dent to consciousness while as transcendence it is a phenomenon of
consciousness, or whether while being a phenomenon of conscious-
ness it has this transcendence only in consciousness-i.e., to speak
with Husserl, this transcendence can be attributed to space only as a
being-characteristic immanent to consciousness. This question
seems to us to be completely aporetical.
Yet it could be asked whether this aporia is merely an expression
of a more fundamental question, and whether the position from
which the aporia stems is inappropriate for a solution because its
way of posing the problem comes too late. For the subject who raises
86. See Husserl, CM, 64; concerning Hartmann (1), pp. 152 ff.
172 Lived Space
the question of the reality of space-and at the same time places the
reality of space in question-has already acted in space, has under-
stood himself in terms of space, not primarily in the naive certitude
of his existence, on which his "natural realism" is founded, but
rather in a more fundamental mode and manner wherein the
certitude of and belief in reality have not yet come into play. In an
originary sense one cannot be "certain" of space-and it is only
when one becomes certain of space that it can be "real" or "tran-
scendent." It is only when space-consciousness, with its beliefs and
theses, attributes certitude to space that there can be such a thing as
asserting, doubting, contesting and defending its reality, or "prov-
ing" and "deducing" its transcendence.
Thus each undertaking of the latter kind-which begins with the
question at the level of the problem of the consciousness of space,
both in realistic and idealistic terms-is left hanging in the air. In
both there is a lack of a secure foundation for an appropriate posing
of the question. Modifying the concepts of consciousness and of the
subject, which stem in both positions from traditionally distinct
styles of metaphysics, does not lead to any advancement of the
problem of space as long as the subject is not sought where he has his
originary relationship to space-in his corporeity. Only a complete
phenomenological illumination of this relationship and the return to
its constitutive strata offers a hope of grasping the question of the
reality of space as such, as a pressing question of a subject who, by
virtue of his reflective capacity, is compelled to ask it. Because of
this very capacity, he must self-critically trace the forgotten primor-
dial questions and pose himself the problem that unavoidably
confronts him as a corporeal being.
PART TWO
MATHEMATICAL SPACE
Introductory Remarks
With the discussion of mathematical space, the investigation
proceeds toward a new and truly complex realm of problems. The
previous analyses began with the directly given comportment of the
corporeal subject in lived space. We must immediately point out that
in contrast, that which is here delimited as the "given," in the sense
of geometric entities that comprise the new region of phenomena,
distinguishes itself from any objectivity given directly and "in
person." lt is conceivable only in a specific mode of access adhering
to strict methodological requirements. In any case, even such a
scientific-methodological approach toward a sphere of objectivity,
though fundamentally distinct from all immediate orientation to-
ward the world as given "in person," does present a genuine--
although genuinely mediated-mode of comportment of the subject.
The modes of the subject's performance contain the conditions for
constructing mathematical theory. Only they can clarify what is
meant by the expression "geometrical existence" found in the
terminology of positive science. They also permit phenomenological
investigation of the specifics of the geometrical mode of inquiry, of
the sense of geometric being and the unique ontological meaning of
mathematical space.
The method, therefore, requires an analytic and all inclusive
clarification. We must not only clearly distinguish the characteris-
tics of this method with regard to its axiomatic-deductive compo-
nents, but also bring into view as a whole those performances of the
subject that are constitutive of geometrization as such. Thus the task
of the investigation is to differentiate between the various kinds of
intuition, as sensory and symbolic, pictorial and significative intu-
ition-between various types of thinking as activities of abstraction
with specific forms of idealization, generalization and formalization.
The ontological meaning of what exists geometrically emerges only
gradually; at first it remains covered over by a network of provi-
sional activities belonging to the clarification of the entire meaning
of geometrical being, which thus ought not to be overlooked here.
Hence it is unavoidable that the following investigation is drawn
into the field of history. No special demonstration is required to
show that the mathematical or mathematizing consciousness
changes and is historically conditioned. That its objectivity is ideal,
that it is constituted as timelessly valid, does not abolish the
historicity of scientific consciousness; rather, it indicates the dis-
tinctive problem with respect to its constituting activities. In this
work the activities cannot be traced in their specifics. Their ques-
174
Introductory Remarks
175
tions are general questions concerned with the constitution of
meaning, and hence cannot be limited to the geometrical domain; to
deal with them would require a comprehensive special investiga-
tion. Nevertheless, it is to be maintained that contemporary work in
mathematics has its motivation and aim in historical roots reaching
far into the mathematics of antiquity. A phenomenological descrip-
tion of mathematics cannot avoid the historical tracing of such
activities. Their historical dimension cannot be overlooked, since
they belong to a region of phenomena, the scientific region, that is
historically constituted in these activities. The phenomenologically
oriented retrogressive questioning is not valid for the historical
changes of scientific facts comprising the history of science, but
rather for the origins of the meaning of scientific entities founded in
the achievements of meaning by mathematical consciousness. To
explcate and clarify these achievements, to reactivate the bestowal
of meaning sedimented in the course of history was, it is worth
mentioning, sketched out by Husserl in a late work as the task of an
"intentional history" of geometry. It is worthwhile pursuing this task
even within a modest framework. Here the concept of geometric
existence is to be clarified phenomenologically along the rough lines
of its structure of meaning.a7
87. Husserl's "intentional-historical" aspect must be understood in a way
that would allow phenomenological retrogressive inquiries to reach not only
to the foundations of sense of the scientific edifice of geometry, but also
deeper, toward the primitive levels of sense constitution; it ought to uncover
the "primordial materials" residing "in the pre- scientific lifeworld" (["Die
Frage nach dem Ursprung der Geometrie als Intentionalhistorisches Prob-
lem"], p. 219; ff. the foreword by E. Fink and the variant text in the Crisis
volume, Appendix III, pp. 365-86).
In accordance with this Husserlian idea of history, our previous research
concerning lived space may easily appear as a partial fulfillment of his
program. An extensive exposition of Husserl's conception of intentional
"history" would lead us too far afield. We stress once again that what has
been accomplished with the three spaces has validated neither "historical"
nor "archaeological" points of view (in the Husserlian sense of founding
stratum and building up of sense mentioned above). Attuned space is
neither a primordial foundation for the spaces of action and intuition, nor
are the latter in the former, i.e., "based" in it. Rather, we had todo with the
three spaces as different modes in which the one actually lived space is
present for a corporeal subject correlated to them. Genuine questions of
sense-founding first emerge for us in the following pages, with the transition
from the natural space of objects to mathematical space.
SECTION ONE
Preliminary Phenomenological
Observations
Chapter One
Space as a Thematic Object of
Consciousness
1. The Space of Intuition as a Limit Case of Lived Spatiality
With the space of intuition, the constitution of the lived space is
complete. Corporeity finds in its intuitive comportment a demon-
stration that as a lived body, it is at the same time transcended
toward its form of existence as a consciousness, in the sense of a
nexus of intentional and uniquely objectifying acts. The intention-
ality of the space-positing consciousness, which is always and
irrevocably the consciousness of a corporeal subject, includes the
ambivalence of a spatial existence as well as the breadth of the
science of the comprehension of space. It conditions the double
mode of appearance of space. This results in the subject's self-
experience as a corporeal being encompassed by space and, at the
same time, as a conscious being who has space unbridgeably "over
against" himself insofar as he makes space into an object. His
"position" in relationship to the intuited objectivities in space
manifests the limits of his corporeal comportment in this space: he
is no longer among things but has moved completely to the periph-
ery of space. The lived body no longer properly constitutes a center,
encompassed on all sides by space, but becomes completely
176
Space as a Thematic Object of Consciousness 177
excentric. It has a wodd of things only as re-presented, which, in
their being "over against," reveal their pure object-characteristics.
As a pure object-space, the space of intuition acquires all those
marks characterizing it as a limit case of lived space in the sense
already suggested: it is the space of a lived body insofar as its
fullness is intuited sensibly and constituted corporeally. Yet it
contains at the same time determinations incomprehensible solely
from the understanding of the subject's corporeity. The space of
intuition did not turn out to be a self-ctmtained part of a spatial
whole that could be thought as composed of the sum of such parts.
Rather, it appeared to be the manner and mode in which the one
objective space is already there for a corporeal subject. The
co-presence of this space also appeared to be constitutive for
sensible-corporeal phenomena, given "in person," "in the flesh."
Thus the space of intuition ceased to be merely the space of a
corporeal comportment, the space of a lived body in a situation.
From the very origin the comportment of the corporeal subject in
space bears the mark of a different kind of possession of space. The
establishment of space as an objectivating positing which provides
the basis for the judgement that space "is" such and such, first makes
space predicable, a subject of valid propositions. This establishment
presents a unique, although neither obvious nor necessarily trans-
parent, achievement of the objectivating consciousness. It estab-
lishes the condition that enables the spatial subject to know and to
designate himself; it makes the awareness and knowledge of the
concrete situation of my lived body in space possible in the first
place.
At the same time, in arder to be able to be claimed as mine, this
establishment would have to justify itself befare each similar space-
positing achievement through the other consciousness, were it not
for the fact that it is the positing of existence of "the" space. It is
neither by tacit arrangement nor by explicit agreement that the thesis
of "the" space means an identical space for all corporeal subjects.
My consciousness of space is mine on the basis of the existential
unity of my consciousness with my lived body. As signifying, as
objectivating, it simultaneously stands for the other consciousness
and indeed for "any" other. Yet in this case it appears in a mode of
universality that even transcends intersubjectivity. While for the
latter the differences between unanimity, majority, and minority
have not yet become completely meaningless, the concern here is
with a form of universality in which any numerical moment is
completely abolished. It is basically this universality that designated
178 Mathematical Space
the "pure" consciousness of the rnuch contested "consciousness as
such." This is by no rneans conceivable without the inclusion of
corporeal conditions and facticity and, indeed, it is only in conjunc-
tion with thern as a structure of factual consciousness itself that it
can be grasped. It rnay not be thinkable asan independent structure,
floating, soto speak, above the individual beings, yet it can becorne
cornprehensible purely functionally in terrns of its specific inten-
tional activities as a corporeal, individual being. Nevertheless, even
so conceived, pure consciousness presents a givenness that cannot
be discredited. Deterrninations of its own can be phenornenologi-
cally brought to light in contrast to factual, individual conscious-
ness, as well asto all intersubjective universal consciousness. It has
its history in its specific universality as the identically functioning
intentionality of "all" corporeal subjects. Moreover, the totality of
subjects that it subsumes rnay require delirnitation against possible
and entirely different consciousness-structures. Y et all of this should
not rnitigate against the application of a concept which, though
rnetaphysically laden, nevertheless appears unavoidable for an ap-
propriate explication and phenornenological analysis of the problern
to be unfolded subsequently.
With this universal consciousness-in its ontic unity with a lived
body on the one hand and, on the other, in its own specific
functionality-possibilities are opened in principie for the corporeal
subject. As already shown, the extrication of the one objective space
already underlies each of his corporeally concrete rnodes of corn-
portrnent and is co-deterrninant even for every structure of lived
spatiality. The expressively attuned and the circurnspectively acting
subject determines his space no less than the "theoretical," rnerely
intuiting subject on the basis of a previously accornplished positing
of "the" space; and the way traced by us above is more akin to a
circle that incorporates its end and its beginning. What allowed us to
speak of the "levels" of lived spatiality was rnerely the increasing
degree of openness and clarity with which the one objective space
carne to consciousness (thetically) for the cornportrnent of a partic-
ular subject himself in a given dornain of his activity. And the space
of intuition rnay be designated as a lirnit phenornenon to the extent
that it reveals deterrninations that could not have been grasped
without the assurnption of objective space in the phenornenological
description of corporeal cornportrnent in the space of intuition.
In the phenornenal dornain of lived spatiality, however, objective
space plays only the role of an inactual horizon of consciousness. It
is "there" for an experiencing, acting, and sensibly perceiving
Space as a Thernatic Object of Consciousness 179
subject; it is "co-given" without any independent intentional glance
being oriented toward it frorn arnong these world-attitudes. Where
such an intention occurs, there appears a new way of regarding space
and a new view of the space-problern. The change from the mere
co-presence of objective space to its full thetic-actual givenness
indicates the fundamental division between a corporeally engaged
subject and theorizing subject, between corporeal functions and
categorial achievernent as capacities of variously constituted and
predelineated forrnations in the essence of subjectivity.
Thus the one objective space becomes open for the many possi-
bilities of topological and metric deterrninations. Its subsumption
under a science whose laws are of a particular type, constitutes one
of the major problems of applied geometry. Viewed from such a
science one would have to discuss the question as to the provenance
of the clairn concerning the euclidean nature of the space of
intuition. From the standpoint of pure geometry, which is the sole
dornain of discussion in this investigation, the euclidean determi-
nation of measure will turn out to be one among many, having no
privileged position within the framework of formal science. Our task
in this part will be to show how the constitution of geometric
manifolds is to be understood on the basis of the comprehension of
space previously discussed.
2. The Topological Structure of the Space of Objects
Objective space functions as the substratum of mathematical
determinations without itself being a mathematical space. The space
positionally assumed in natural consciousness as "real" space is
primarily without metrics. It is still on the hither side of metrics and
as such has a series of structural properties that are in no wise
specifically mathematical.
The uniqueness of this space, in contrast to all lived spatiality,
appears most clearly in its isotopy and homogeneity. In its absolute
oppositionality to the subject it is without any firm directional
determination, without specific valence of place. In complete
counterdistinction to the spatial cosmos of antiquity, articulated in
accordance with center and periphery, the space of our contempo-
rary consciousness of objects does not accord any value to its
locations: it is apure system of "places."
Its homogeneity immediately raises the question of its topological
structure. It was already mentioned that the concepts of endlessness,
boundlessness, and infinity are topologically ambiguous. In the
180 Mathematical Space
phenomenological section of the previous part an exact differentia-
tion among these concepts was deliberately avoided. Whether space
is open-endless, or whether its infinity is to be understood as
inauthentic, in the sense of boundless yet closed connectedness, was
still of no importance for the problems encountered there. Moreover,
such a distinction is lacking in a consciousness of space for which
the one homogeneous space is indeed "there" but only in the mode
of co-given presence. Phenomenologically speaking, such co-given-
ness appeared only as a going-over-further of space beyond the
intuited horizon.
lt is only where space becomes thematic in objective conscious-
ness that topological differences first become relevant. It is illumi-
nating that when the homogeneous space of objects is conceived as
a manifold of empty places, it can in principie be open-infinite, at
least as a noncontradictory possibility of thought. This is also valid
if for sorne reason the corporeal world should be regarded as finite.
During our preceding investigation of the empty space, we men-
tioned that, and precisely how, the space of this topological structure
dominates modern consciousness, or at least its scientific-physical-
istic thinking. For the newly emerging problem, certain trends of
thought must now be mentioned briefly in arder to show that the
infinity of space is not merely assumed; rather, an attempt will be
made to offer an insight into this structure of infinity on the basis of
phenomenologically sufficient grounds.
Husserl had attempted to conceptualize this infinity of space by
analyzing determinations that are characteristic of the singular space
of intuition.
88
His notion of the infinity of space, which delimits
phenomenologically the sense of Kantian idea, ought to be under-
standable from motivational interconnections originating with the
immediate experience of things. If the experience of a thing is in
principie inadequately given by a limited appearance, then, accord-
ing to Husserl, the complete givenness of the thing is prefigured as
an idea, as a system of endless processes of continuous appearing in
whose constant transition one and the same thing "is more precisely
and never 'otherwise' continuously-harmoniously" determined.
This continuous perception of a thing, determined "more precisely
as infinite on all sides," also includes all spatial adumbrations of the
thing as "movable in infinitum; thus with it we grasp the idea of
space, or more precisely, the idea of its infinity. For Husserl this idea
is found precisely in the factually inadequate grasp of the complete
88. See Husserl, Ideen 1, 143-149.
Space as a Thematic Object of Consciousness 181
objective unity of a thing, a grasp that is only possible as limited and
thus must be constantly thought as open-infinite.
A critical pursuit of this argument reveals its inadequacy for the
problem under discussion, precisely because of the unnoticed am-
biguity in the Husserlian concept of infinity. It is quite thinkable that
each continuous transition through the various appearances of a
thing is in truth not determinable as open-endless, but rather as
reiteration. In contrast to genuine infinity, it may contain merely an
endless repeatability of a closed and thus in principie completable
series of thing- appearances. This in no wise contradicts the "idea"
of the adequacy of a thing as such.
89
Similar reservations were already voiced with respect to the thesis
of Scheler (pp. 157 ff.). In fact, it seems that despite the substantive
difference of their starting points, the argument of Husserl and
Scheler formally have the same conceptual structure: both maintain
a factually limited possibility which, nevertheless, is actually inex-
haustible. It is of the type in which further possibilities always
remain continuously "open," i.e., "open" in the sense of the genuine
infinity. Yet it remains unnoticed that the openness here meant as a
possibility-either in grasping things or in corporeal movement-
and as a possibility in the only proper sense, namely, that of the
unfulfillability in principie of a claim on the part of the finite,
corporeal being, cannot lead to a univocal decision between two
completely different, and in this difference "open," possibilities.
That is, there is a residuum of undecidable possibilities stemming
from the underlying interconnections of phenomena. It is precisely
because all factual experience of a thing terminates in a finite series
of appearances, and precisely because all factual corporeal move-
ments inevitably constitute only a limited space, that they leave
89. A precise phenomenological analysis could, as a matter of fact, show
that Husserl's infinite-continuous succession of partial intentions and
respectively of their syntheses of fulfillment is only a closed-endless
succession. This is related to the fact that each thing is in principie movable
in infinitum. Yet for the complete display of its adumbrations, one requires
rotations; translations alone are not sufficient. At this point, Husserl has not
availed himself of one of his earlier insights. Already in the third chapter of
his Sixth of the Logische Untersuchungen we find that in the exposition of
the characteristic features of "fulfillment" within the more extensive class of
identification, there is an insight into an "aimless" identification in the
perception of the thing without a progressive approximation toward an
epistemic goal.
182 Mathematica1 Space
open in principie the possibility for two basically distinct topolog-
ical structures at large.
Another solution is offered by Carnap. Beginning with the limited
space of intuition, he regards the open infinity of the homogeneous
space of objects as a guarantee for the "required" extension of the
space of intuition.
90
This conception is indefensible insofar as it fails
to do justice to the fundamental fact that the topological character of
space as a whole does not depend on the arbitrary choice of the
space-apprehending consciousness. Space as it is in the natural
consciousness of objects neither allows a choice of connexus nor
does it leave this connexus in darkness so that something definitive
could first be made of it by a specific kind of requirement. Rather, the
open infinity of space is there and is unavoidably present in the
already characterized manner in which it is "given."
Nevertheless, Carnap's claim can be maintained phenomenologi-
cally. It is based on the justifiable insight that the infinite structure
of space-even if seen topologically as one among various possibil-
ities-is in a specific way a distinctive structure. If we revert to the
phenomenon of horizon in the space of intuition (pp. 111 ff.), we
notice that the clearly co-present "continuation" of space in the
consciousness of the intuited limit, which first constitutes the
horizon as horizon, attains its complete and genuine meaning only
with the co-posited open infinity. This side and the hither side of the
limit has here a univocal sense-fulfillment. The univocity is most
clearly graspable with the phenomenal relationships of ordering.
These relationships-such as the relation "between"-are univo-
cally continuable only under the idea of a genuine infinity. That
given three phenomenal points, one lies between the other two,
constitutes a univocal arder in the space of intuition. Yet in a
continuation conceived to extend beyond the horizon, the arder
would lose its univocity if it were to be assumed that the line
connecting them was closed. In other words, space as an open-
infinite space is distinguished from any other topological connection
in that "in the large" it is conceived to be topologically the same as
90. R. Carnap's concept of the space of intuition does not correspond
strictly to ours insofar as he do es not accept this space at the outset solely as
a space of the sensible concrete thing of intuition; rather-by appealing to
Husserl's seeing of essences-he sees it in terms of geometric formations and
relationships. This aspect is still lacking in our investigation, since for us
this problem falls within that of geometric space. Nevertheless, what is
decisive here is solely the required character of extension as such.
Space as a Thematic Object of Consciousness 183
in "the small," i.e., accessible as a segrnent of space to concrete
intuition.
91
Only in this way can it be present to natural conscious-
ness-indeed not as a requirement but rather as a factual and
sensible continuation of the limited space of intuition. Thus the
sense of that which was understood as a co-possession of the
homogeneous space of objects in the individual space of intuition is
now displayed from another side. If on the basis of the phenomena
discussed in the first part the space of intuition had to be conceived
in such a way that the one homogeneous space is there at the outset
for the corporeal subject, then what has just been presented shows
the reverse side of this state of affairs: the latter space can only be
represented in the former because its connexus is congruent in all its
parts with the space of intuition.
91. Subsequently, the all-pervasive planarity of space will emerge as its
topologically predominant characteristic (see pp. 248 ff. of this work).
Chapter Two
Basic Trends of Mathematization
1. Morphological and Mathematical Determinations of the W orld
of Things
Philosophical thought about geometry reveals an ongoing contro-
versy consisting of two main questions: one is concerned with the
object of geometry and the other deals with the "origin." With
respect to the first, the conception that geometry is a science of space
must be countered by the fact that geometry deals with formations in
space but not with space itself. The question of origin touches upon
the genuine metaphysical problems of geometry; within the tradi-
tion, apriorism and empiricism present two extreme attempts at a
solution.
We are primarily concerned with finding an appropriate access to
both problem areas from what we have already elaborated.
In arder to describe the world of objects, even in the space of
intuition, the subject is led to a series of characteristics which, in
accordance with their sense, belong to two completely distinct
modes of observation. Characterizations such as notched, jagged,
egg-shaped, umbelliform, etc., stand among others: triangular, circu-
lar, spherical, cubical, etc., which are basically distinct from the
first, owing to the fact that in contrast to the first, they are not purely
morphological characteristics.
Such morphological characteristics-while unavoidable and even
sufficient within the domains of knowledge that belong to them as,
for example, the descriptive sciences-are merely "vague" determi-
nations; they present themselves as "fluid" and admit of various
displacements in the spheres of their application.
92
As descriptive,
92. Concerning the concepts of vagueness and fluidity, see E. Husserl,
Logische Untersuchungen, II, Prolegomena, 21, and Ideen I, 74. Let it be
recalled that the Husserlian concept of vagueness is not identical in meaning
with imprecision, but only with not-exactness. Its positive determination
results from the above context. See also O. Becker (1), pp. 398-401.
184
Basic Trends of Mathematization
185
universal determinations they are extensional concepts and, as such,
they maintain a horizon that is open in principie for possible real
objects falling under them. Each singular object strengthens anew
their sense of being a universal concept and the number of such
objects is essential for the "degree" of universality attributable to
these concepts. They are, furthermore, vague concepts because they
are indeterminate in relation to other features belonging to the
individuals that are apprehended under them. For example, objects
with otherwise completely heterogeneous features are found under
the concept of egg-shaped. The contemporaneity of these features is
at a given time present only empirically and thus is merely factual,
constituting a contingent aggregation but not a necessary intercon-
nection. The morphological concept of the egg-shaped cannot be
deciphered from these remaining features. The complex of these
features, although present, remains outside of the range of the
extensional concept and is not apprehended with this concept; it
remains completely undetermined by it. At bottom, this state of
affairs designates the language of abstraction, which purportedly
leads to such concepts. On the one hand, it means an extrication of
a singular feature and a "leaving," a "looking away from" all
remaining features; on the other hand, it includes an "ascending" to
concepts of a specific level of universality which grasp more objects
in common. This fact is most likely to be visualized in the image of
a pyramid of concepts, wherein the greater the extent of the
concepts, the lesser is their content. Such a morphological concept,
as an extensional concept, has its specification "under" itself but not
"in" itself.
lt is otherwise with the determinations mentioned in the second
place in the enumeration above. In accordance with their specific
meaning, they are not morphological but-in a sense to be discussed
in more detail-mathematical determinations. This does not contra-
dict their morphological use in the space of intuition. After all, in the
pre-reflective consciousness of sensory intuition the categorical
difference of both types that emerges here does not at all come into
view. That this is the case is nevertheless nota proof for the generic
correspondence of all their form-properties or of a common
subsumption under the superimposed concept of "morphological
feature." The fact is that this common and consciously nondif-
ferentiated usage of both form-types in the space of intuition must be
understood from the side of the subject: with the use of these
non-morphological features in the space of intuition, the subject
turns out to be a participant in a differently construed domain of
186 Mathematical Space
concepts and universality. These features are concerned with
characterizations of the sensible contents by a being capable of
comprehending types of forms different from those that are consti-
tutive for the descriptive sciences, namely the forms of mathematics.
In sensible intuitive comportment, they are primarily and for the
most part meant morphologically and are employed as extensional
concepts. Insofar as they retain this meaning, the properties belong-
ing specifically to morphological features are also valid for them. In
this sense what is valid for the features of the umbelliform or of the
jagged is correspondingly valid for the features of a triangle or a
sphere. What is characteristic of the latter is that they are not
exhausted in this meaning. These new determinations are not
depleted by their morphological sense when such determinations
are used for classification.
The fundamental difference between both kinds of determinations
consists in the following: with those that are to be recognized as
mathematical "universal concepts," the complex of all the addi-
tional features is no longer completely indeterminate. Such concepts
as, for example, triangularity, sphericality, etc., are essentially so
constituted that with them are posited other determinations that
have a necessary conceptual interconnection. In distinction to the
morphological features, we have here a simultaneity of properties
that are not present merely factually, to be apprehended through
enumeration and thus conceived conjunctively only in this mode;
rather, a specific 1ogica1 structure shows up: the absence of one of its
remaining detreminations leads to the disruption of the entire
contexture and the negation of it yields a logical contradiction. Any
of such new concepts is airead y a conceptual structure with a
completely determined logical structure; in this case it not only
contains other determinations under itself in the same way as a
generic concept would "subordinate" various species concepts, but
contains them truly in itself. Hence it is possible to conclude the
entirety of the structure from each of its members.
Their validity, therefore, cannot be strengthened by any new
empirical discoveries, nor can their universality be confirmed by
such discoveries. The concept of a square does not become more
universal simply because more objects are discovered possessing the
property of squareness. Either a present or imagined multitude of
things displays all of the properties that logically correspond to the
concept of a square-orthogonality and equality of length of the
bisectors and of the diagonals-or the properties are not given. In the
latter case, it would not mean that the concept of a square has
Basic Trends of Mathematization 187
become senseless or useless and has no "universal character";
rather, it would mean that those objects are not square. This reveals
the fundamental difference between morphological and pure math-
ematical determinations: the number of things lending fulfillment to
the intuition of a mathematical "feature" is completely irrelevant for
its meaning-content and does not dictate its universal validity. For
mathematical concepts there is no "degree" of universality. The
universality of mathematical concepts is not equivalent to that of the
universality of genera.
Another related fact, not adequately considered befare, is that in
the geometrical domain there is no strictly generic construction of its
objects. Such a construction requires a graduated series of increasing
or decreasing generality and specificity in such a way that the more
specific contains the universal in a specific manner. The latter can be
obtained by excluding the specific differences of the species, so that
it appears to contain a lesser content in contrast to the specific and,
conversely, the specification of the genus occurs in such a way that
through the addition of specific differences there emerge sub-species
richer in content.
In contrast, the relationships in the geometric domain are essen-
tially different. The wholly elementary mode of defining sorne for-
mations does indeed occasionally masquerade as a generic-logical
structure. The rectangle, defined as as parallelogram with right an-
gles, or the circle, grasped as a closed, curved line with a constant
degree of curvature, etc., are apparently present as species under
certain higher forms of genera.
93
Yet it must be observed that what
appears here in analogy to differentia specifica has an entirely dif-
ferent form of "addition" than in the above case. Strictly speaking, it
does not consist of an additive incrementation of something new, not
yet co-present in the "genus," but merely in a specification of what
was already present in the universal-and indeed, present variably.
Let B = [B
1
, B
2
, . Bn] be posited for a geometric concept B with
93. Husserl also seems to maintain this generic scheme for geometry. As a
paradigm, he mentions the triangle as an (eidetic) singularity under the
highest genus polygon (Ideen 1, 12). His polemic against Locke follows
from this assumption (LU, Investigation 2, 11). This is all the more
astounding in that Husserl had already established the important distinction
between generalization and formalization (Ideen 1, 13). At best, this is
explainable by the fact that in elementary geometry-to which Husserl refers
exclusively and only occasionally-this differentiation is hardly formed. (In
addition, see the following presentation.)
188 Mathematical Space
determinations B
1
, B
2
, Bn, with a resulting special concept B' =
[B't, B
2
, Bn; B'
1
, B'
2
, B'm] stemming from B, not through the
fact that additional determinations B't, B'
2
B'm are simply added
to the first, independently of those of B (i = l ... n); rather, the
geometrical specification takes place in such a manner that one each
of B' (j = l ... m) proceeds from an establishment of a demarcation
of any of B. This specification can continue as long as new variations
are to be encountered under the obtained new determinations B', B"k
etc. A most extreme particularization of this form results in the con-
cept whose determinations are established completely through con-
stants; for geometry this means through determinate numbers.94
What constitutes a "special case" in geometry is thus not the
endowing of a universal with additional, new determinations, but
rather specification of the universal itself by specific establishment
of its various determining parts. At the same time, a geometric
formation is capable of mathematical specifications-contains spe-
cial cases of itself-insofar as it contains variables. Here the more
universal concept is at the same time the richer in content. The
universal triangle T = [s
1
, s
2
, s
3
; a, J3, -y] is no poorer in features than
the right-angled triangle T' = [s
1
, s
2
, s
3
; a, J3, -y; a = 90] and the
right-angled equilateral triangle T" = St. s
2
, s
3
; a, J3, -y, 'Y = 90; s
1
=
s
2
]. Both are special cases of the universal triangle not because they
are contained determinations that, in contrast to the composition
of the universal triangle, possessed something new not yet previs-
ioned in it, but rather because the previsioned is actualized in them
94. Specific formations can also arise from E = [a ; b] through various
specifications: E = [a; b, a>b], ore = [a; b, a = b] ""'e = [a]. As a simple
example for the theory of conic sections, we can use the ellipse in the
2 2
presentation of its equation of center \ + Y
2
= 1, whereby the half axes
a b
are variable independently of one another. (The concept of variables is
obviously used in the logical and not the analytic geometrical sense, since it
only has the functional variables x and y.) By specifying a = b we get a circle
with radius a and finally the specific circle, such as a = 4, no longer
containing any variables.
In addition, it seems that the ongoing controversy concerning the "uni-
versal triangle" of Locke, eliciting up to the presenta plethora of logical and
psychological discussions, rests, to say the least, on a misunderstanding of
the universal character of geometrical objects. An additional problem
certainly plays a role here: the relationship between the geometric object
"itself" and its "figure" at the level of signs. (See pp. 196 ff.)
Basic Trends of Mathematization
189
in a specific form. Thus while completely deviating from the work of
the classificatory sciences, the mathematician sees at the same time
a scientific criterion of value of his efforts in that they correspond to
the aim of the greatest possible universality. He does not investigate
singular and separate formation in arder to reach "abstractively" to
higher ones; rather, he seeks the possibility of universal structures,
and what he has proven for them, he has eo ipso also proven for the
special cases. Conversely, where the work is initiated with individ-
ual formations, the universalization does not proceed so that the
specific characteristics are neglected, but rather so that they are
taken as variables, or at least, no use is made of their specificity.
Such a construction of the geometrical domain of forms is obvi-
ously different from that in the classificatory sciences. To be sure,
within elementary synthetic geometry its uniqueness is difficult to
grasp. The positing of constants of variables, constituting the appro-
priate process of specification, is a logical operation. It first appears
in the analytically pursued geometry and thus first attains its
meaningful expression in the functional equation. There is no
correspondence with the pictorial-symbolic since here it is easy for
morphological elements to intrude into the conception. (Thus the
structural difference between conic sections in elementary geomet-
rical presentation and their projective application in functional
equations!) A beautiful proof that the division of the formations in
the old synthetic geometry were similar to genera and species is
found in the fact that in the geometry of antiquity, the theorem of the
sum of angles was proven separately for the different "species" of
triangle. The process was sensible and necessary as long as the
insight was lacking that none of the triangles investigated are
something "specific," but rather are mathematical special cases of
the universal triangle.
The autonomy of mathematical determinations, in contrast to the
empirical world of things, appears when we note that the former are
completely free from the role of being a mere feature of things and
are grasped purely in themselves. Indeed, they must be grasped in
this mode if their specifically mathematical character is to come into
view at all. All further properties belonging to the characteristic
"spherical" will be studied not with spherical things, but with the
spheroid as such; what allows us to recognize any of the properties
and relationships appropriate to the concept of a triangle is not the
triangularly bounded objects, but the triangle itself. At this point let
us not enter into a closer analysis of the "sphere" and the "triangle"
themselves; let us first take their meanings as they are understood
190 Mathematical Space
prior to any reflective individual analysis by objective conscious-
ness, and let us note that here it is a question of a new and unique
domain of objectivity. It gains its independence not because the
objects of the space of intuition are completely sorted out in it in
terms of the presence of these new determinations; rather, they
manifestly bear the certainty of their existence in themselves by
virtue of their factual and relational interconnections which, more-
over, cannot even be exactly realized in the objects of intuition.
This means that the access to this domain of objects must be at the
outset other than the access to the morphological domain. The
mathematical concepts cannot be abstractively obtained from the
intuited objects in the same sense as the morphological concepts.
This reveals the difficult problem of their origin (pp. 191 ff.).
If we briefly characterize further, purely phenomenologically,
their mode of givenness for a nonreflective consciousness, then the
new objects constitute an ideal interconnection. This will mean that
the essence of these objects, and of the relationships obtaining
between them, lies apart from the awareness of the "flowing" time.
Concepts such as development, becoming, change, and even those of
constancy and continuity, become nonsensical in this domain. The
new objects are what they are: trans-temporal and valid trans-
temporally. Furthermore, they are apprehended in a specific mode of
universality. The triangle, the circle, the sphere, are distinct from
any individual interconnections. While being identical, they are
singular and, at the same time, are not momentary individuals in a
here and now; rather, they are ideal singularities untouched by
principium individuationis. The most extreme specification to be
attained in this new domain of objectivity is the ideal singularity.
However, the above delimited relationships do not yet designate
the mathematical concepts as mathematical. The entire domain of
ideal aspects is not exhausted by the mathematical domain. It must
accordingly be asked wherein lies the uniqueness of the former.
2. The Problem of Mathematical Ideation
Types of entities other than bearers of expression, ready-to-hand
things, and objectively intuited entities have a different mode of
corporeal comportment as their correlate on the side of the subject.
The interconnection of things constitutive of the types of things in
lived space, as well as co-constitutive of their relevant spatiality,
provides a foundation for a special unity of sense constituted in
specific apperceptive activities of a corporeal being. These activities
Basic Trends of Mathematization 191
are such that they form both the lived body and the thing in their
specific mode of being thus.
This correlative conjunction of lived body and lived world became
obvious with the reflectively oriented analysis of corporeal move-
ment in space. The correlation appeared as a result of a retrogressive
observation of the immediate comportment toward the world with-
out itself being given within and in the accomplishment of the latter.
In unreflective comportment toward the world, things are available
as pre-given in another mode of presence. The general and encom-
passing mode of presence of things in lived space is one of direct
self-givenness. For something to be given directly means that in its
apprehension there are no mediate moments of conception-such as
being a sign oran image for another. To mean something as itself, in
a strictly phenomenological sense, means to intend it not in a merely
signitive mode as happens, for example, in a merely verbal under-
standing or in all symbolic thinking; rather, it means that it is intuited
as "itself" present and has its fulfillment of meaning in intentions of
a "seeing" kind, where various differentiations become immediately
significant for the latter. In addition, each objectivity of lived space
bears a character of self-givenness "in the flesh". This presence "in
person," "in the flesh," means literally the mode of givenness of an
object in its confrontation with functioning corporeity.
The domain of objectivity now coming into view lacks first of all
this character of presence "in the flesh." The givens of mathematics
have no direct relationship to corporeity. The many equivocations of
the concept of intuition run the risk of enlarging the difficulties of
comprehending the domain under consideration here. They contain,
on the one hand, the roots of the numerous and all too extensive
attempts to found geometry empirically. Such attempts think that
they can obtain the geometrical formations from the sensible-
intuitive factors-i.e., from what is given "in person" of the lived
space-through abstraction from the morphological determinations
in the sense analyzed above. On the other hand, the equations also
give rise to those apriorisms that purport to guarantee the objects of
geometry by presenting them in an unconditioned "pure" and
purely "spiritual" vision, having nothing in common with the
corporeal achievements of the subject and thus completely tran-
scending all sensory functions as such.
For the phenomena to be analyzed here Husserl has coined a
concept, that of ideation, and demonstrated its fruitfulness as a
mediation between two extremes. As a specific case of abstraction, it
participates in the usual aphairesis. lt disregards all peculiarities
192 Mathematical Space
present in the individual objects in their particular hic et nunc. At
the same time it is essentially different. It does not merely bring forth
something common in a given multitude of objects; rather, on the
basis of perception "in the flesh"-presentation and imagination of
an exemplifying individual object-it intuits the universality of an
essence.
Accordingly, universal essentialities are obtained intuitively "on"
an individual object of sensory intuition. This seems to be similar to
the notion that we are concerned with extricating something con-
tained in the object and first apprehended as an empirical feature
befare the ideational regard can really orient itself toward it. Mean-
while, even if the ideation is enacted primarily on the basis of
"direct" sensory intuition, it already presupposes the presence of
this very essence as one and the same, in contrast to a multitude of
objects of the same type and form. With this ideation a new act-
characteristic appears in which a new type of objectivity is consti-
tuted. Indeed, the acts remain founded in sensory intuition, yet they
manifest their new type of intentionality in such a way that the
founding acts are not included in this objectivity. It contains new
determinations that are not found with the objects of the founding
acts.
Ideational domains of objectivity are constituted with a specific
character of universality. Their objects, as essential universalities
and ideal objects, are indifferent not only to the number of empirical
cases, but also to the possibility of their empirical realization at all.
Nevertheles, the ideationally constituted object appears immedi-
ately andas self-given; yet this self is no longer a self simpliciter and
"in person," but a categorial one. The Husserlian concept of catego-
rial intuition includes quite appropriately the layering of two
intentionalities: the sensible and the intelligible. Yet it neither
merges into unitary act-quality (as is the case with the various partial
intentions in "simple" perception, i.e., within one act-level), nor is
it composed of two independent and closed series of acts ("sensi-
bility" and "understanding"); rather, it presents a unique complex of
acts of sensible and categorial moments which, in accordance with
their phenomenal composition, justify the extension of the concept
of intuition to include categorical structures.
95
We are well aware
that ideations, with respect to their phenomenological structure, are
95. Concerning the concept of ideation see E. Husserl, LU, Investigation 6,
48-52; Ideen 1, 70 ff.; and CM, 34. Concerning Categorial intuition see
E. Husserl, LI, Investigation 6, particularly Section 2, Chapter 6.
Basic Trends of Mathematization 193
very special and very simple intentionalities within the total domain
of categorical intuitions. We must explicitly recall that the latter can
be understood and justified phenomenologically without contradic-
tion only in the precise sense of the much contested Husserlian
concept of "seeing essences." Furthermore, it is primarily with the
help of this categorical intuition that the reference to sensory
intuition as the "source" of geometry attains a more precise sense.
And it is this mediation that first offers an explication of the most
discussed concept of geometrical intuition. A closer exposition of
this will follow (pp. 253 ff.).
But befare this, it must be pointed out that something ideationally
intuited is not merely not sensible; rather, above all, the ideational
abstraction as such is not yet mathematical ideation. Its designation
requires additional characterizations. Thus the latter distinguishes
itself from other kinds of ideational performances essentially
through a unique moment of idealization-and this must be taken in
the specifically mathematical sense. What is grasped ideationally is
not only conceived in its pure essence as a trans-temporally valid
mode of universality, but beyond this is also grasped exactly. A
spherical object is never exactly spherical, and yet the discovery of
this inexactness suggests that its form is already addressed in
immediate intuition as spherical and thus at the outset is apper-
ceived as an ideal and exact form-structure.
This state of affairs is a sign not only of mathematical ideation as
distinct from any other grasping of essence, but also as the specific
complexity of the act structures designated by the concept as
mathematical ideation. To a great extent phenomenological analysis
allows what constitutes these structures to escape. O. Becker first
clearly realized the problem emerging here, with all of the difficul-
ties, and characterized mathematical ideation as an ideational ab-
straction with a certain transgression of boundaries with appropriate
limits of structures. Through the contraction of topological nets this
abstraction "sharply" lays out the mathematical formations as points
of the mathematical continuum.
96
What ultimately constitutes the
ideality of the geometric formation in a precise sense can be grasped
only from the background of the mathematical continuum. The
genuine problem of mathematical ideation has its objective correlate
in nothing other than the mathematical continuum. Yet its problem
is no longer geometrical: it falls, with all its impenetrability under
96. O. Becker (1), Section I, Part 1. Concerning the problem of point and
continuum, see pp. 296 ff. of the present work.
194 Mathematical Space
the domain of mathematical analysis. Since at this point we are not
yet engaged in the investigation proper to it, it is premature to talk
about it. Here the primary concern is to indicate the specific
difficulties appearing with the question of the origin of geometry.
These difficulties increase if it is noted that the mathematical
ideation, taken in a precise sense, is still inadequate to allow the
pecularities of geometric objectivity to emerge. What appears solely
as given in this ideation are certain discrete forms, elementary
geometric formations limited in number which not only do not
comprise the whole of geometric objectivity but are not even
comprehended as specifically geometric, i.e., as functions subsumed
under an axiomatic-deductive science. The specifically mathemat-
ical sense-bestowing-for example of an exact eidetically appre-
hended triangle "as" geometrical formation with all of its constitu-
tive moments of meaning-in its logical formal nature is not
something that merely accrues to, but rather simply transcends
ideation. It first determines the constructive character of mathemat-
ical objectivity, grounds its typically mathematical interconnec-
tions, and makes it demonstrable within the framework of a rational
theory (pp. 211 ff.).
3. Symbolic Intuition (Pictorial Symbolism)
The relationships and differences between morphological and
mathematical objects must be conceived from still another view,
which is usually neglected in the treatment of the theme that is here
at issue. A closer look leads to noteworthy insights.
The ideational intuition of geometrical formations turned out to be
founded in sensory intuition. The former thus contains clues point-
ing back to the corporeal functioning of the subject engaged in
geometry. This does not exhaust the relationship of geometric
objectivity to corporeity. It is rather characteristic and most remark-
able for this objectivity that its full complexity of meaning unfolds
only gradually and can be mastered on the way through a mediation
of something else that is not geometry. This something else is its
"presentation" in sorne medium of the sensible intuitive world. To
attain valid propositions concerning geometrical formations we
require a figure or-to speak with reservation-a "picture" of it in
which it "shows" itself.
97
This self-showing in another is not
97. For the context under discussion here it is irrelevant whether in
presentation we are concerned with a figural ("pictorially symbolic") or
Basic Trends of Mathematization 195
something external in the object, something merely added; rather, it
will turn out to be a constitutively essential aspect of the geometrical
and mathematical domains as such. All that is mathematical is such
only insofar as it is productive in another.
98
However, the geomet-
rical object thereby becomes highly complex. No longer appre-
hended in direct intuition "in person" and yet not apprehensible in
any other manner except in something else given "in person,"
apprehensible as "itself" meant in what is given without itself being
in it-on the one hand, this complicated state of affairs contains the
entire problem of the geometrical sense of being and, on the other, it
presents the preeminent question of geometrical intuition as well as
the intuitability of geometry.
For a closer analysis of this intuitability there is an extant concept
which, while enhancing the distinction between the figural triangle
and the triangle itself, at the same time obliterates the genuine
structure of this distinction. It is the concept of representation. The
figure ought to "represent" the properly and geometrically meant:
this means that the representandum has already attained a clear
separation from the represented, and yet that the former is appre-
hended in the latter as "itself" in its independent meaning. But this
condition can be the basis for a misunderstanding. It would lead to
the notion that the mathematician takes the sketch, instead of the
meant ideal triangle, and studies in it what he wishes to discover in
the triangle itself. But this would mean that he would only obtain
information concerning the triangle that is offered in the sensibly
perceived formation. Thus he would not reach geometrical knowl-
edge as such, but only a morphological description of a figure-
which like any morphological figure would not be exact. Y et
admittedly, the imprecisely sketched triangle is sufficient for exact
knowledge of its geometrical properties; even an elliptical figure
may be of service in understanding the properties of a circle.
Although requiring the visual figure, this understanding does not
rest in it, but rather in an appropriate consciousness of the meaning
"signitive-symbolic" presentations. (Concerning the distinction see 4
below.) Here we are concerned only with the presentational function as
such.
98. Related to this is the constructive character of the geometric formation
(pp. 216 ff.). "To be producible" here is a general and precursory delimita-
tion of the constructive moment in geometry. In the history of mathematics
its meaning has been variously transformed; here it can be traced only
gradually.
196 Mathematical Space
of the ideal states of affairs. In geometrical work we do not "live" in
a pictorial or symbolic consciousness, but entirely in the conscious-
ness of the thing itself. It is the articulating reflexive analysis that
first separates both; the separation must be assumed in arder to
reveal the mediating character of this new kind of self-givenness.
The term of symbolic intuition, and primarily pictorial-symbolic
intuition-in contrast to signitive symbolism-is thus comprehensi-
ble not from the mathematical intuition, but from phenomenological
reflection.
99
Subsequently we shall use the concept of symbolization
with the explicit restriction that in geometrical work we do not mean
the co-given consciousness of the pictorial mediation of the object
and of the symbol as "symbol"; rather, the sale characteristic of the
intention of consciousness of the meant ideal object itself. It is this
fact that first makes possible the meaning of the figure as a pictorial-
symbolic figure for the geometric state of affairs; symbolic intuition,
in the sense meant here, is not motivated by the observation of an
"abstract'' painting such as a triangular composition.
The difficulty here lies in the difference between the ideal and the
spatially real ("sketched") formation or, seen from the side of the
subject, between the ideal consciousness of meaning, wherein the
thing "itself" is constituted on the basis of categorial intuition, and
the simultaneous and essentially necessary co-positing of its sign in
sensory intuition. Self-givenness and givenness "in the flesh" are
sundered here in a different way than in a mere sensory intuition
simpliciter. The ideal object itself is also immediately given; the ray
of ideating intuitionality encounters it directly without traversing
signs as the medium of meaning "of" the object. However, the
geometric object is not self-given in sensory intuition, but rather in
a categorically intuiting intention. Yet to be fulfilled, confirmed, and
shown to be valid, it requires the complementary activity of sensory
intuition functioning as a symbolic intuition in relation to the meant
objectivity.
The discrepancy between the ideal geometric formation itself and
its intuited figure underlies the already suggested problem of the
universal triangle in Locke. Locke starts with the fact that a geomet-
ric proof is always accomplished with a specific triangle-more
precisely, with a figure at the level of a sign which can obviously
represent only a completely determinate triangle ( containing no
variables-yet it was concluded that the proof is valid for any
99. The concepts of pictorial and signitive symbolisms used here follow E.
Husserl, Ideen l, 43.
Basic Trends of Mathematization 197
triangle. Locke hints at this conclusion in passing, claiming that the
proof at the outset relates to the "general idea of a triangle" and
hence is valid per se. Thus for him the ineradicable paradox appears
between the fact that this universal triangle must have all the
specific determinations, and yet at the same time could not ha ve any
of them. At the same time he maintains that although this triangle is
completely devoid of specific determinations, it unifies all particular
triangles in itself.
Husserl criticizes this "absurdity" with the remark that the con-
tents of the various "kinds" of triangles are subsumed by Locke
under the concept of a triangle.1
0
A phenomenologically adequate
solution of the problem is not found in Husserl. lt was not even
found with Husserl's presuppositions in which the classificatory
scheme of genera and species was applied to self-evident, geometric
formations. As a matter of fact, it would be absurd to prove the
theorem of the sum of angles in terms of a specific triangle and then
to conclude that this was valid for any triangle or for "the" triangle,
if the latter, as genus proximum, were a product of abstraction from
"all" triangles and if, thereby, none of the particularities of the
formation represented in the figure accrue to it. Yet it must be
observed that here no inference takes place-apart, of course, from
the conclusions immanent in the proof. No inference is present from
a specific triangle to other, less specified triangles and finally to a
universal triangle as such. This is already an indication that there is
a special relationship between the figural triangle established as a
sketch in all of its determinate parts and the triangle. This relation-
ship must be considered if the concept of representation is here to
provide more than a way out of a perplexity.
101
A representation,
after all, does not only have the characteristic of representing
something trans-sensory in something sensory; rather, the problem is
to show that and how all specific triangles can be represented in a
100. J. Locke, Vol. II, Book IV, Chapter 7, Section 9; see also Chapter 1. For
Husserl's critique of Locke LI, lnvestigation 2, 11.
101. C. Holder (1) advocates the view that geometric proof always con-
tains an analogical or completely inductive inference (p. 12). This opinion is
obviously attained regressively; the proof is valid for all triangles, "there-
fore" such inferences "must" participate. Yet this cannot be maintained
phenomenologically in the factual process of proving; it does not even
accomplish what is required to save the universal validity of geometric
proof. A formal-logical solution of the problem is offered by the work of E.
W. Beth.
198 Mathematical Space
singular figure! Such a representation, if it were to arder the most
diverse triangles under genera and species, would be impossible-
justas impossible as it would be to say that all musical instruments,
be they only the string, the percussion, and wind instruments (and
under the latter the wood-wind and brass instruments) can be
"represented" by a single sketch of a bassoon.
In truth, even the starting point of the discussion requires a
correction. It cannot be denied that the sketched figure is always
specific, i.e., it has fixed sides and angles; what can be debated,
however, is that it stands for a specific ideal triangle, comprising, as
it were, its congruent mapping. The fact not only is that the actual
size of the figure is of no interest at any moment but that in geometric
work itself its factual determination of size is not actually present to
consciousness; this should warn us against any attempt to import
sorne meaning of depiction into the concept of representation meant
here. In its project the figure is taken at the outset in its universal
dimensions as a, b, e, and <t, 3, 'Y While being sketched factually in
quite determnate sizes in the figure, they are nevertheless meant as
variables. We do not abstract from sizes for a successful process of
proof or for an inference to a "genus triangle"; rather, during the
proof we make no use of the possibility of their constancy in arder
to assure the greatest number of variables for the result, in such a
way that a variable contains possibilities of determination while
encompassing all possible specific cases. Thus the process does not
lead from a sensible, singular figure by way of abstraction and
generalization to a universal triangle lacking all particularities.
Rather, it is categorical intuition as such that first constitutes the
"triangle" as an ideal triangle itself and, indeed, in a figure with
variably conceived determining parts-"in sorne figure or another"
-and it is the latter that plays a role, though an unavoidable role, in
the process of proof. In arder to see the figure as factual and
determined entirely by the length and position of chalk marks, a
specific-and, indeed, seen from the point of view of the geometric
process-a contrived shift in viewpoint is required. The Lockean
problem arises only when the contrived viewpoint is assumed at the
beginning; here one overlooks the difference between what the figure
is as a formation given in direct intuition and what it means from the
outset in the project as a symbolic-intuitive presentation of some-
thing non-sensible. The Lockean conception of the problem fails to.
pay attention to the specifically universal character of the "univer-
sal" triangle and fails to see the essence and function of pictorial-
symbolic representation in geometry.
Basic Trends of Mathematization 199
The concept of pictorial-symbolic intuition lends preliminary
understanding to the meaning of the statement about intuitability in
geometry. This must be clarified later. First of all, we agree that the
concept of intuitability, in case it is not yet specifically delimited,
must be taken in the sense of pictorial-symbolic intuitability. While
reserving more precise determinations for subsequent discussion,
here it is possible to establish sorne provisional outlines: to appre-
hend a geometric formation means to orient oneself to the thing
"itself" in a pure consciousness of its meaning, and to be able to
correlate certain sensible-intuitive elements to it and to its ideation-
ally apprehended properties in such a way that (1) the correlation of
the ideal state of affairs and the pictorial content is reversible so that
not only is there a correspondence between the state of affairs and
the pictorial content representing it, but also the pictorial content,
insofar as it is grasped as merely sensible in the primary intention,
can nevertheless be constantly apprehended through an appropriate
reorientation as symbolic for the relevant states of affairs; and
moreover, that (2) the possibility must be given in principie for
apprehending the pictorial symbols univocally as pictures of possi-
ble morphological forms of the world of things given "in the flesh"
and meant as depictions.
While the possibility indicated for (1) is as valid for a model
conception-such as in non-euclidean geometry-that is distinct
from the pictorial-symbolic intuition as it is for the signitive sym-
bolism of analysis, the requirement mentioned in (2) is specific to
the geometrical intuitability meant here. It will be shown that the
latter concept is delimited in a way that precisely encompasses
euclidean geometry. In accordance with the delimitation indicated
in (2), the circle, for example, of euclidean geometry is in its
presentation geometrically intuitable in the above specified sense as
a closed and equally curved line. In contrast, the straight line of
elliptical geometry is not intuitable in such a presentation of a circle.
As "straight," the picture of the closed line calls forth, in direct
intuitive consciousness, an experience of conflict. The circle func-
tions in this case no longer as a pictorial symbol but rather as its
"mode1",
102
since phenomenologically it is classified and consti-
tuted differently. It becomes clear that the required correspondence,
co-present latently in the pictorial-symbolic intuition, is exploded.
Thus if non-euclidean geometry is to be presentable in any pictorial
102. Concerning the question of the model conception, see pp. 278 ff. of
this work.
200 Mathematical Space
way, it requires a model-but there is no model of euclidean
geometry.
4. Signitive Symbolization of Geometry
Categorical and pictorial-symbolic intuition designate only two of
the kinds of intuition here in question; however, they do not
characterize the entire domain of acts that constitute geometric
objects. What comprises their geometric nature ultimately remains
foreign to ideational and idealizing abstraction as well as to the
forms of intuition mentioned. What will be intuited ideationally,
what will be established pictorial-symbolically in a real medium, are
merely specific elementary formations, along with a few geometric
relationships in them. Thus with the right triangle one can only
intuit the content of the triangular asymmetry; yet that the precise
relationship between sides and hypotenuse is quadratic, as ex-
pressed in the Pythagorean theorem, is not at all intuitable either in
ideational abstraction or in symbolic intuition. The typically math-
ematical character of this domain of objectivity does not come fully
into view within the modes of intuition observed to date.
The elementary geometric formations considered heretofore are
not fundamental formations in the sense of being objects for foun-
dational investigation. The point of departure of foundational inves-
tigation is comparable to ours in only one respect, insofar as its
questions can begin on the basis of an airead y pre-given and evolved
individual science. Yet it diverges essentially from our investigation.
In contrast to ours, it does not explicitly include the subject in its
investigations and thus remains oriented purely objectively. Indeed,
the foundational investigation of the geometrical domain of objec-
tivity is also for its part made into an object of reflection. Thus
Hilbert demands: "We must make the concept of the specifically
mathematical proof itself into an object of research, just as the
philosopher criticizes reason itself."
103
Yet the investigations of
axiomatic theory and of the theory of proofs are maintained within
the context of the discipline of mathematical thinking and are
accomplished with mathematical and logical means. While founda-
tional investigation discovers the logical structure of a domain,
projecting such structures anew through new axiomatic beginnings,
formalisms, and calculus, our interrogation aims at the analysis of
those constitutive activities of consciousness from which the giving
103. D. Hilbert (1), p. 415.
Basic Trends of Mathematization 201
of mathematical meaning is first discerned at all. Our investigation
faces a problem that is prior to any axiomatic researches; it encom-
passes the latter not in their results, but in the point of departure of
their question. It is directed to this question: what appears, beyond
the geometric region as such, as a necessity for the aim of mathe-
matical researches into foundations? How must it be constituted in
its ontic structure and how must it be articulated ontologically so
that it would become comprehensible as a region requiring special
researches?
If the geometric objects were only those of intuition, whether
ideational or categorical, then no mathematical investigation into
their foundations would be required. Since geometry is concerned
not only with direct but also with categorical and symbolic intuition,
it means that the mathematical aspect is present only conditionally
in intuitive orientation. The clearly discernable discrepancy be-
tween the meaning-consciousness of states of affairs themselves, on
the one hand, and their sensibly perceptual presentation, on the
other, led to the problem of symbolization in the region of elemen-
tary geometry. For a long time it was taken as a pictorial represen-
tation of individual geometric formations, which are meant when the
natural consciousness ascribes "intuitability" to geometry. Yet here
the symbolic process in geometry is first established only in its
factual and historical origin. What constitutes the progress from the
old synthetic geometry to the modern analytic geometry, as seen
act-phenomenologically, has its grounds in the dissolution of
pictorial-symbolic intuition by a signitive symbolization, by a rep-
resentation of the meant in signs that are fundamentally distinct
from the geometrical "picture."
Phenomenologically speaking, sign-symbolism in geometry
presents a new and most unique problem. It is far removed from
being a mere and supplementary establishment of what can be
present in pictorial symbolism; on the contrary, the sign points toan
entirely new kind of sense-bestowing in the geometrical states of
affairs themselves. We are not thinking primarily of modern formal-
isms in metamathematics, but of the earliest form of algebraization of
geometry anticipated by Descartes. What is genuinely algebraic in
geometry, what allows it to become "analytical" geometry, cannot be
investigated here in sufficient detail, and to a great extent must
escape phenomenological-descriptive efforts. First of all, such
efforts would have to observe separately all those acts and their
multifarious syntheses that constitute number, arithmetical laws,
algebraic operations, etc.; this would require comprehensive specific
202 Mathematical Space
investigations. In their stead we can suggest Husserl's Philosophy of
Arithmetic. At the outset our investigation must be limited: it must
show that with the synthetic and the pictorial-symbolic, on the one
hand, and with the analytical and sign-symbolic, on the other, we are
not merely speaking two distinct geometrical languages, whose
meaning can be accurately and mutually translated; rather with the
transition to analytical presentation, there appears something fun-
damentally new. That, on the one hand, a curved line at the sign
level represents, in pictorial symbolism, a mathematical curve of a
higher arder, and that on the other hand, a combination of signs F (x,
y) = O "presents" the "same" curve, except now not pictorial-
symbolically but in signitive symbolization, indicates that this
linguistic parallelism fundamentally hides the deepest riddle inher-
ent in the latter kind of pres8ntation. The special science of geometry
employs the concept of isomorphy here in arder to designate the
structural equivalence between the geometrical region of objectivity
and the algebraic region of operations; for its aims it is sufficient to
remain with the conception of a "mapping" of one domain onto the
other.
104
Meanwhile, the problem we have indicated has not been solved.
Even after a successful elucidation of all constitutive achievements
of pure mathematics, another question would remain open: how is it
possible to posit an arithmetical formation for a spatial one? To
signify points of space with numbers-how is this fundamental act
of analytic thinking to be conceived, a thinking that was still remate
to the geometry of antiquity? Obviously the numbers of a series are
present as unique individuals such that no two are alike. In contrast,
the space of geometry is given as a homogeneous manifold in which
none of its points have geometric characteristics distinguishable
from another. It seems that through the coordination of numbers and
points, the latter become fully individuated; while in elementary
geometry none of them were distinguishable, now each seems to be
distinguished.
If the sense of that isomorphy is to be conceived, the coordination
between number and space must, as a matter of fact, remain
completely incomprehensible. But what then assures the possibility
of such coordination is not the circumstance that the world of
numbers "consists" of nothing but individual numbers; rather, it is
that we have to do with a domain in terms of its generative
104. K. Reiderneister (3) develops sorne of the rnain problerns of geornetry
under this aspect of isornorphy.
Basic Trends of Mathematization 203
principies and its being built up purely constructiveJy from a few
basic conceptual operations. In this domain the specificity of indi-
vidual numbers is in no wise abolished. Rather, they are submitted
toan all-pervasive order and lawfulness such that this domain can
validly be called homogeneous under the opera ti ve aspect. From this
vantage point the question of the application of number to geometry
begins to make sense.
With respect to spatial formations, however, this application is
less than self-evident. That the extended spatial formations which
are distinguishable and indefinitely multifarious in position, size,
and forro, should be mastered collectively by the constructive arder
of the numerical domain is not an immediately obvious fact. The
mathematical notion of "mapping," in its statics of coordination,
hides the fact that we are not dealing here with the mutual relation-
ship of two equally complete domains, but with a new methodoJog-
icaJ postulate for geometry. The geometrical formation "ought" no
longer to be taken as a rigid and immobile pictorial-intuitive figure;
rather, it ought to be thought as emergent from pure number
sequences in such a way that the arithmeticallaw of their coordina-
tion determines the geometric forro. And conversely, the properties
that can be brought to light pictorially in a geometrical formation,
such as direction and curvature, must turn out to be characteristics
of these number sequences.
Thus it seemed that initially the domain for this new geometrical
mode of conception was circumscribed very narrowly. Even
Descartes still had to exclude the transcendent curve from his
consideration. What testifies to the inexorability of the basic Cartes-
ian conception is the fact that while it was established for a limited
geometric domain, and subsequently became a problem for addi-
tional geometric formations, this new beginning was not discarded
as inadequate. To subsume geometric formations collectively under
the number domain and, in a final analysis, not even to admit
anything as "geometrical" if it cannot be expressed in pure numer-
ical relationships or (to use a later term) be dealt with "algorithmeti-
cally," is to show that this mode still dominates geometry as a
methodological postulate (pp. 219 ff.). Infinitesimal calculus and
differential geometry methodologically signify merely the conse-
quent fulfillment of the Cartesian demand.
With this, however, geometric objectivity not only appears here in
a transformed symbolic forro; rather, what is more decisive is that its
analytic "conception" introduces at the same time a complete
factual modification of meaning of what is geometrical. If geometry
204 Mathematical Space
has become preerninently analytic geornetry, then, for example, the
Pythagorean theorern concerning right triangles is no longer valid for
the measure of line segments, but solely for the algebraic state of
affairs: what is given here is "pure quadratic form." Number rernains
predominant in its own intrinsic rules, the equation. and its "form,"
which are not sensibly formative but purely algebraic.
105
Purely
quadratic formas the square of the "distance" shows a characteristic
shift of meaning-giving whereby a geometric state of affairs appears
to be nothing but an interpretation of an algebraic state of affairs.
And this shift is not merely one of psychological thought or a matter
of mere attentiveness, but is graspable in the very sense of the
ontological meaning of 'geometry.
With this sense-modification of what is geometric there suddenly
open up vast possibilities for the broadening of geometry. For exam-
ple, in accordance with its purely algebraic meaning, the new concept
no longer includes a limitation to two or three dimensions. Only the
degree of the equation determines its geometric meaning, while the
number of its members is the matter of choice. Yet at the same time
this choice determines the possibility of multi-dimensional geome-
tries. Furthermore, the "existence" of various geometric manifolds,
the metric structure of entire mathematical spaces, depends on the
pure algebraic form of a c ~ r t a i n equation, the fundamental tensor.
Finally, an algebraic criterion is posited over a geometric meaning.
Yet it is not abolished as geometrical. Rather, in its algebraic modi-
fication, it is expanded to encompass all elementary geometric for-
mations and states of affairs. What is of primary importance in this
context is that the transition from the elementary synthetic geometry
to analytic geometry is based on a profound transformation in the
domain of symbolizing acts. These are completely novel modes of
consciousness in which the "analytical" is constituted-although in
this novelty they are not limited to the mathematical domain. Yet they
are mostly significant in this domain, since its radically changing
methods are deterrninative for a completely altered conception of the
scientific character of geometry.
The older geometry remained essentially caught up in pictorial-
105. Certainly, even such a form as a combination of sign-symbols
possesses its own sensory shape anda "pictoriality" proper to it. However,
the intuition corresponding to it is to be strictly distinguished from the
signitive-symbolic intuition to be analyzed here (pp. 209 ff.). In the present
context the primary concern is with the moment of meaning; the signs
function as symbols "for" something that they themselves are not.
Basic Trends of Mathematization 205
symbolic representations of its formations. In its apprehension of
geometric existence it was still exclusively bound to pictorial means
for the constructive acquisition of its objects. The analytically
operating geometry requires the dissolution of the pictorial element
through the sign-symbol. With the concept of constructability it
modifies at the same time the existential meaning of the
geometrical. In terms of act-phenomenology, we are concerned with
a sequence of levels of meaning-giving with a very complex
structure. It constitutes mediations of various kinds with respect to
the apprehended object which, while easily achievable factually,
are describable phenomenologically only incompletely and with
difficulty.
The pictorially intuited figure presents the geometric formation
itself, even if in a limited sense of appearing as itself through the
mediation of the figure; yet in and with the figure, the ideational
intention truly touches the meant, so that the figural-pictorial
medium points directly to the subject matter itself, without hin-
drance and without deflection of the view. In contrast, signitive
symbolization contains various media of meaning with distinguish-
able "indices of refraction" behind each other. Thus in accordance
with the motivation of the act, either their seriality is surveyed
completely or the intention remains at one stratum and there
independently constitutes an interim "self." Thus, for example, any
discussion of a circle, where it is meant in a pictorial-symbolic
intuition, is different in its meaning from a circle when it is spoken
of as a "circle x
2
+ y
2
= a
2
." Indeed, in the latter case a specific
equation admits coordination to the circle as also meant with the
signs, yet it is no longer represented in the sense of the term used
above. What is here "represented" in the sign, in strict analogy to the
above usage, is not the circle but the functional equation of the
circle. The circle in its turn is mediately intended by the equation.
The nominal positing of the circle in the first case has, correspond-
ing to the act, not the positing of "the circle x
2
+ y
2
= a
2
," but
rather "the functional equation of the circle x
2
+ y
2
= a
2
." Here
even in naive execution of the mathematical intention-and not, as
in case of the pictorial-symbolic representation, first in the reflective
attitude-there is a clear co-consciousness of the positing of the
functional equation for the geometric formations. In other words, in
signitive symbolization, the various apprehensions of the sign artic-
ulate themselves as "standing for," as "pointing beyond" in the
objectively oriented course of acts; or else the acts attach themselves
to the sign. Both states of affairs, again differentiated within them-
206 Mathematical Space
selves and to be discussed shortly, are typical for the en tire phenom-
enal complex of signitive symbolization. Included is the notion that
the relationship between sign and signified-in contrast to that
between picture and its object-is no longer direct. This also implies
that signitive symbolism is not prescribed by the meant object. While
the pictorial symbol remains bound to the symbolized, thus retain-
ing the "picture" character and, if need be, allowing inexactness at
least within the permissible morphological limits, the sign-symbol
contains the possibility in principie of a free choice. It requires that
the relationship between the sign and the signified for which it
is to stand rnust be established through agreement. The signi-
tive-symbolic rneaning intentions involve specific acts of choice of
the positing of signs, conventions of their meaning-giving, and
requirements of univocity of their usage. Here an entire sequence of
distinct modes of intentionality determines a multi-leveled media-
tion of the meant object. The ray of intention first encounters the
object on a variously disrupted path, through distinct and phenom-
enologically heterogeneous levels of meaning, before it appears "in"
the sign as that which is rneant throughout.
This multi-leveledness offers in its turn specific and phenomeno-
logically singular possibilities. It was already mentioned that in
signitive symbolism, in contrast to the pictorial, the apprehension of
the signas "sign for" is inserted into or joined onto the very course
of the objective acts themselves. The first case is appropriate where
something geornetric is explicitly meant in the signs, for example in
analytic geometry in a narrower sense. There the signs present
functional equations, and as such they already "really" reveal
geometric formations, relationships, and processes, i.e., the latter are
explicitly meant in the former. Yet nothing stands in the way of
either the striking out or, at the very outset, the neglect of the
specifically geometrical meaning of such symbols. The intentions
could rernain with the equations and functions "thernselves" and
could terminate with the sense-bestowal of the functions as func-
tions, as algebraic equations.
With this restriction to the purely algebraic theory of equations, the
defacto relationship between something spatial and its sign is dis-
solved. The establishment of signs is simply a matter of choice. In this
case the mathematician can speak of a "geometrical interpretation"
of algebra. This mode of observation deviates from a genuinely geo-
metric observation; it is not motivated geometrically, but rather is
primarily algebraic in kind. An additional although necessarily re-
quired intention reinstates the relationship to the spatial formation.
Basic Trends of Mathematization 207
This is not to say that this "merely" interpretative observation is
unimportant for geometry-indeed, with it an entirely new sense of
the geometric as such is being delimited. In the first place, such
interpretations yield an extension of the concept of geometry. Its
various branches owe their existence only to a preceding algebraic
theory, an existence that could not have been constituted without
the algebraically mediated intentions, i.e., without using the alge-
braic process. Thus, for example, skew field geometry is based
originally on an algebraic property, namely on the non-commutative
multiplication of its underlying algebraic number field.
Conversely, the mathematical concept of interpretation, which
was only able to assume its full extent and develop its extreme
consequences after Hilbert's meta-mathematics, yields a conception
of the geometric that has been fundamentally transformed-not only
in contrast to antiquity, but also to modern mathematics. What is
present in it as geometry is no longer a solely geometric inter-
pretation of algebra, but in the last analysis more of a geometrically
interpreted calculus, a geometrically interpreted formalized theory.
Furthermore, this means that we are concerned with a conception of
the geometrical whose original motivation is no longer the question
of how an objectivity, conceived primarily as spatial, can be
mastered mathematically, but rather is dictated from an entirely
different and obviously opposite question: namely, what can a
system of signs offer to a purely operatively established geometry if
one attributes geometrical meaning to it. This conditional statement
brings to expression a completely unbound possibility. Hilbert's
well known witticism that for a, a, A, etc, one must be able to say at
any time "tables," "chairs," shoes," "tankard," characterizes the
complete unboundedness of interpretation of modern signitive
systems.
106
Yet remarkably, there appeared at the same time a completely
new kind of meaning within the signitive domain itself. To use
Hilbert's citation, while the old symbolism was founded in an
106. A question might be raised whether we are dealing here at all with
geometry or whether we should allow it to be circumscribed in the
traditional sense against mere "geometric interpretations." This cannot be
decided here. In modified form this question is closer to that of mathemat-
ical "spaces" and we shall cometo them subsequently (pp. 239 ff.). It has not
been claimed here that geometry belongs today to operative mathematics.
We are only attempting to trace possibilities that can result even for
geometry, in the broadest sense, with the process of signitive symbolization.
208 Mathematical Space
ingenious insight that for "tables," "chairs," "tankards," one can
also say a, a, A, modern operative mathematics stands the
relationship between sign and signified, so to speak, on its head.
The originary achievement of mathematical symbolization-which
produced its signs "from" an objectivity through which something
pictorial and imaginable was represented, not meant in an imageless
way-appears all too obvious today. The signs of such images were
meaningful and were akin to the original symbols as a presentation
of an objective sense. But in calculus the sign has become removed
from the signified to such a degree that not only is the intention
toward the signified, the external and independent objectivity
actually omitted, but its omission is explicitly demanded. If
traditional symbolism stressed the representative function of signs,
thus meaning them "merely" as signs for something, in modern
science the sign is taken for itself. Indeed, in principie it has an
open horizon of possible interpretations, capable of subsuming the
geometrical at will; yet it is primarily signified as itself and for itself
in independent "meaning," and only as a sign, without signifying
anything that is not a sign. Having no posited semantic background,
the "sense" of the sign here is no longer representative, but exhausts
itself in being purely operational. The sale concern is not what these
signs mean, what they signify, but that and how they can be
managed.
Such a "formalism" may at first sight present an exact opposite of
the older sign syrnbolism, yet considered phenornenologically, it is
only the end of a path already initiated by the latter. The seeming
transformation of the relationship between sign and signified in the
modern theory of signs is in truth nothing else than the most extreme
consequence of what was posited as signitive possibilities in the
original symbolization through signs.
Thus the problem of intuition and intuitability appears once again
from another side. While it could be adequately solved for the
pictorial symbolism of geometry, it seems that it can no longer be
meaningfully posed for a purely signitively treated geometry, and
especially for a purely signitively projected geometry. If the sale
concern is to interpretan "abstract" configuration of signs spatially,
then any possibility of intuitability obviously becomes completely
redundant. Obviously, signitive symbolization is essentially distin-
guished from pictorial symbolization in that it lacks any aspect of
intuition. Thus in the founding of geometry, Hilbert never evokes
intuition. This is all the more remarkable since iti another place he
claims that intuition is actually the most secure source of knowl-
Basic Trends of Mathematization 209
edge, and indeed even in the very foundation of analysis.
1
0
7
Yet
there is also a danger of equivocation here. To understand Hilbert's
expression adequately, we must respect a twofold differentiation: on
the one hand, between the pictorial-symbolic and the signitive-
symbolic, and on the other, between sensible and categorical intu-
ition. That the sign a does not represent a straight line in pictorial
symbolization, but symbolizes it signitively, and that furthermore
the meaning "straight line" does not have a sensible but only
categorical-intuitive fulfillment, is quite obvious. If the concept of
intuition is restricted by the character of sensibility, or at best by
pictoriality, then a geometry done purely signitively obviously lacks
any intuitability.
Nevertheless, a complete exposition of the concept of sign retains
an intuitive moment. What remains ineradicably intuited, and
indeed in the narrower sense of the direct sensory perception "in the
flesh," is the sign itself. This is the case even when it is taken to be
a sign "for" something spatial. However, this moment of sensibility
in sign-giving attains specific accentuation if the sign stands for
itself, if it becomes an object of mathematical consideration. Insofar
as it no longer has a genuine representative function, but is posited
in its own meaning as purely operational, the sensible intuition of
these signs "in person" assumes an eminent mathematical signifi-
cance. In fact, it plays a leading role in the process of proof.1
8
Indeed, the signs of metamathematics are not geometrical forma-
tions, yet whatever their posited meaning may be, they are nonethe-
less spatial formations. Thus no operation with them is nota process
of thought that merely becomes subsequently fixed in a spatial
medium; rather, from the very beginning the operation necessarily
occurs through signs and not by way of the signs. Although chosen
arbitrarily as to their kind, they are not externa! or superfluous to
thought; they are rather constitutive for what occurs in and through
107. D. Hilbert (3), p. 158.
108. K. Reidemeister (1) speaks of a "method of self-clarification and
self-control of an exact structural intuition" in mathematics (p. 201), and
clearly sees that the concept of intuition in geometry is laden with
ambiguities. Yet upon consideration of these ambiguities, his view that
intuition in mathematics has merely the form of a motive, a conjecture, an
overview, ora plausibility (p. 198) cannot be maintained in this summary
form. Even the notion that the interconnection between the modern system
of "intuitable propositional signs" and the space of intuition can be found
solely in that the signs can be "spatially interpreted" (p. 205) does not
adhere to Reidemeister's own previously stated remarks.
210 Mathematical Space
them. Thus at the same time, mathematics possesses in the config-
uration of signs, a means of control of its statements. This is of a
peculiar kind. Owing to the fact that signs are spatial formations-
more precisely, that they are sensible-morphological figures in a real
spatial medium, each logical contradiction in calculation is grasp-
able in direct morphological intuition. This is not merely valid as a
possibility; rather, the direct sensory perception of the signs be-
comes a necessary act of mathematical operations themselves.
lt is crucial that the signs are not just spatial formations in
themselves, but that they stand in a completely determined and
nonarbitrary spatial ordering and are subject to strictly established
rules of operation. Furthermore, it is significant that this spatial
ordering is linear. This does not merely mean that it is progressively
realized and at the same time a temporal arder. Every elementary
geometrical, pictorial-symbolic construction is also temporal; it is
established progressively in time to the extent that elementary
mathematical thinking is already discursive. Yet the finished con-
struction does not divulge the succession of "steps" of construc-
tion-in any case, it can be regressively seen "through" in its
becoming, but cannot be seen "into" progressively. lts elements
stand in a temporal arder of succession only in the process of
construction; they are spatial because they have constitutive func-
tion for an objectivity, exclusively in terms of the requirements of the
subject matter.
In contrast, the relationships in signitive symbolism are quite
different. The reiteration of an identical sign ata "subsequent" place
first manifests the temporal continuity of process. However, as a
place for something sensibly perceivable it appears as "later" not
only because it is merely "other" spatially, but also because beyond
the manner of spatial ordering of signs, it is so constructed that it
presents time in an irreversible direction of earlier and later. This is
the fundamental meaning of the linearity of their arrangement. They
are so constructed that the signs simultaneously reflect the arder of
thought. Here the logical "sequence" is apprehended distinctly and
immediately as a spatial sequence. The linearity of the configuration
of signs manifests the discursive structure of mathematical thought
in a sensible-intuitive way.
The present observations are not designed to overlook the enigma
attendant upon the unique intertwining of acts of "pure" logical
inference with those of direct perception "in the flesh." Obviously
the ordering of signs, purely as such, can be neither contradictory
nor noncontradictory. How is it possible that a mere combination of
Basic Trends of Mathematization 211
sensible and sense-empty signs in space can signify correct and
incorrect inferences? Behind this question is a more fundamental
problem concerning the conditions for the possibility of any appear-
ance of something non-sensible in something sensible. This is of a
speculative nature and escapes phenomenological analysis and
description. The latter can only show how such an appearing
factually occurs in a process of a correlative and implicative admix-
ture of act-intentionalities. At least it may illuminate the sense of the
"concrete" that Hilbert has attributed to metamathematical signs. By
disregarding any depictive function, the sign, precisely when it has
become devoid of images and has become posited in its own right,
has reached a new intuitability for itself and thus a capacity for
surpassing the achievements of pictorial symbolism for the construc-
tion of new mathematical structures.
5. The Constructive Character of Geometric Objectivity.
Geometry as a Demonstrative Science
The great discovery of the Greeks was that mathematical objectiv-
ity is capable of and above all requires scientific, indeed rigorous
methodological access.to9 This mathematical method is more accu-
rately characterized as an axiomatic-deductive method. And geom-
etry becomes knowledge to be proven-thus it appears as an
achievement related to a thinking of a very specific structure.
This structure is usually designated as logical and even mathema-
ticians call it "purely" logical. "Logical" thinking in mathematics
ought not to be taken in the narrower sense, as if it proceeded in
accordance with classical rules of inference. Syllogisms can hardly
be said to appear in mathematics. Various attempts to formulate
mathematical proofs in syllogistic form have yielded very little and
frequently were based on illusory syllogisms.
110
It is otherwise with
109. Since we are not engaged in historical investigations, we shall not
enter into the question whether mathamatics prior to the Greeks was given
an incipiently scientific treatment. More explicit discussion is found in O.
Becker (4)-(7), K.v. Fritz, O. Neugebauer, H. Scholz, and A. Speiser. The
investigations of v. Fritz place a special emphasis on the extent to which
Greek mathematics already has exact inceptions, so that Becker (5) could say
of Euclid that he "appears more like an equalizing compiler than a creator in
his own right" (p. 20).
110. Instead of designating mathematical thinking as non-syllogistic, it is
more fruitful to designate it positively as "algorithmic." Concerning the
concept of the algorithm, see pp. 219 ff. of this work.
212 Mathematical Space
a relationship of the mathematical modes of inference to the formal
logic of relations. To discuss this relationship would lead us too far
into the researches of mathematical foundations. It must be stressed
once again that our investigation is not designed to engage in
mathematical foundations. Our reflections concerning proofs in
geometry do not comprise a theory of proofs in the sense of special
researches. They must rather be accepted as given; they belong to the
phenomenal region just as elementary geometry do es, and must
therefore be regarded in the same way as the latter. At the same time
it is to be recalled that modern theories are only comprehensible in
light of their historical development, insofar as their inquiries are
motivated by previous conceptions of problems; thus the uniqueness
of their achievement emerges clearly from their own historical
background.
Such a motivational nexus appears in many forms and will be
elucidated under various aspects. The problem of intuition in
particular reappears with regard to the specifically scientific char-
acter of geometry. Now we must investigate the role of intuition in its
distinct types, as they were analyzed above, and the way it functions
in the total configuration of the activities of demonstrative sciences.
Furthermore, the constructive character of the geometrical must be
considered. The concept of construction can be clarified only in
relationship to the changes of its meaning in the course of history.
Finally, the question of the algorithmic structure of mathematical
thought leads to the contemporary foundational investigations, and
in particular, its results must be contrasted to Husserl's charcter-
ization of the algorithm.
With regard to the question of intuition, to be taken once again as
pictorial-symbolic representation, it obviously plays a very re-
stricted role in the process of proof, even in ancient mathematics. K.
Reidemeister views the achievement of Euclid to be precisely a
"transformation of the intuitable into the conceptual."
111
Euclid's
work is characteristic of the striving to exclude as far as possible the
predominant employment of sensible-intuitive experience in an-
cient mathematics and to provide an axiomatic base for it. This
means that for the formation of fundamental principies he selec-
tively employed only a few and obviously unavoidable facts, intu-
itively required. Everything else was acquired purely mentally with
the systematic recourse to these principies leading to the creation of
a strictly rational method. Nonetheless, what justifies our speaking
111. K. Reidemeister (2), p. 51.
Basic Trends of Mathematization
213
of a transformation lays too much stress on the transformation and
gives too litle credit to the role 0f intuition in the founding of
Euclid's geometry. Geometry even befare Euclid was demonstrative
and, as shown today by researches into the history of mathematics,
the need for axiomatics has a long prehistory. Attempts at proofs
appeared first and were followed by exact definitions befare Euclid's
axiomatic theory emerged. It can be said that his axioms are
principies in the Aristotelian sense: they are first in this science,
although not first for us. Due to the lack of axioms of ordering,
Euclid's system of axioms, however, was so construed that in many
cases of demonstration, intuition was introduced not merely "for
illustration," with the aim being a subsequent observation of plau-
sibility, but rather was unavoidable in the very process of demon-
stration.
In the geometry of Euclid it is inappropriate to limit the role of
intuition solely to the establishment of axioms. This view is contro-
versia! for la ter geometry, at least from the aspect of concrete
geometrical work. The methodological characterization of geometry
as a deductive discipline designates an essential trait of this science;
yet this applies in general to its finished structure, so to speak,
without giving adequate account of the factual process of the
geometer in which this structure is first of all constructed. It assumes
that in geometry one proceeds in such a way that all subsequent
results are obtained from pre-given or assumed premises purely by
means of logical inference.1
12
But even with the contemporaries and
followers of Hilbert, the mathematicians do not proceed so mechan-
ically in their work. Rather, in more recent geometry the procedure
is as follows: with a "proposed" statement, i.e., with a statement that
is assumed to be true mathematically, one seeks the presuppositions
required to demonstrate the statement, and quite often the opposite
assumptions turn out to yield a true statement. Thus to a certain
extent the results are here given first while premises are discovered
subsequently. To what extent such assumptions are occasioned by
intuition in geometry cannot be stated in general and must depend in
112. See O. Holder (2), pp. 26-28. L. Nelson (p. 387) also claims that as
soon as the axioms are assumed, then all theorems follow "through the mere
form of inference." In any case, the non-euclidean geometries, with which
Nelson is primarily concerned, present an entirely new aspect of the
problem. They initiate a transformation of the concept of axiom, later
culminating in the work of Hilbert. Concerning the question of axioms and
their meaning, see our subsequent discussions, pp. 216. ff.
214 Mathematical Space
individual cases on the content of what is being assumed. At least
this could hold true when the result (a theorem) is more intuitable
than the premises, which is quite often the case. As a theorem it may
be valid only when it is obtained deductively as a conclusion
through a chain of inferences; in such cases the function of intuition
is not demonstrative but "merely" motivating, although it is relevant
for the technique of demonstration. To limit the significance of
intuition to the acquisition of intuited fundamental principies and to
relegate everything else to "pure" thinking is to misunderstand the
nature and origin of such thinking.
Even for the thought of antiquity, however, it is crucial that
mathematical work must encompass the intuited aspects, to the
extent that such work begins with something intuited and proceeds
with symbolic intuition. Even what is immediately and intuitively
evident in the figure, such as a geometric proposition that is valid for
the states of affairs themselves, must in turn itself be demonstrated.
Even such a simple relationship as triangular inequality is not aided
by having recourse to the evidence of intuition; if it is valid as a
mathematical state of affairs with respect to the triangle, it must be
deduced from something else.
Even if each step in thought is "followed" in intuition, and if
intuition must present and retain what is grasped in each step, the
real concern is not with the intuitive, but with what "follows"
logically. In accordance with its essence, the logical is so constituted
that it does not require any pictorial-symbolic illustration. What is
announced in the latter are at best certain factual and partial results
"leading" ultimately to the conclusion (e.g., auxiliary lines of a
construction). The logical consequence as such escapes pictorial
symbolization. Its essential characteristic, the inferential "then,"
cannot be presented in any pictorial-symbolic way. Only signitive
symbolism can establish it in a sign and reach not only a new kind
of representation of geometric formations but also and ultimately an
exact presentation of geometrical propositions concerning these
formations.
It is demonstrative thought that is responsible primarily for
revealing the ontological meaning of geometric propositions. That
something is geometric is synonymous with the fact that it is
provable; this is valid not only for the meaning of the copula in
geometric propositions concerning states of affairs, but also for
existential propositions. Even in the mathematics of antiquity, the
existence of an elementary geometric formation-e.g., a circle or a
triangle-is not assured simply because it is seen ideationally and
Basic Trends of Mathematization 215
represented pictorial-symbolically; rather, its existence must be
genuinely demonstrated. This identity of being and being proven
constitutes the uniqueness of the geometrical as well as the mathe-
matical sense of being. A geometrical state of affairs does not first
exist and then get demonstrated; rather, it comes into being in the
process of demonstration. It has its being in nothing other than in its
being demonstrated as valid in a process that in its turn is capable of
rational verification employing criteria of validity and invalidity. As
invalid-proven as not valid-mathematical states of affairs are
nothing. They may exist for a while in an individual consciousness
as a psychic construct, but it does not emerge into objective,
mathematical existence. This is not counter to the circumstance that
there are unproven propositions (e.g., the great theorem of Fermat).
That clearly noted exceptionality in the total field of deduction
confirms what is stated, not as an exception would confirm a rule--
mathematics has nothing to do with rules of this kind, and its
complete interconnectedness is not such a rule--but in the way a
single case always disturbs mathematical thinking, as a deviation
from the ideal mathematical science. Such an ideal remains preemin-
ent in its attempts at proofs. The more difficult circumstance seems
to be that mathematics contains undecidable propositions, proposi-
tions for which no proof is attainable concerning their validity or
invalidity. At the same time, this undecidability is of a unique kind.
It is incomparable to any other manner and mode of uncertainty, and
on precise grounds. Here the structure of identity between being and
being proven is displaced. What is demonstrable is not the mathe-
matically posited state of affairs, but its undecidability in principie.
The non-demonstrability of the truth or falsity of such propositions
is itself provable precisely.
The identity asserted is also not contradicted by the fact that
geometry, like any deductive system of propositions, also contains
axioms-fundamental principies that cannot be demonstrated.
Aristotle saw clearly that axioms as such belong to a demonstrative
science. The intent to prove everything in a demonstrative science,
including the fundamental principies, would lead to an infinite
regress and thus never to a science. If a science is to be established
as a demonstrative science at all, it must necessarily proceed from
propositions that are not demonstrated. Moreover, the latter must
also be undemonstrable; if they were demonstrable, then, in accor-
dance with the concept of a strictly demonstrative science, every-
thing that is demonstrable would also have to be actually demon-
strated, and thus the demonstration would have to be given for the
216 Mathematical Space
principies as such. Undemonstrable fundamental principies are a
necessary presupposition of any demonstrative science.
113
Thus a decisive question is: how is demonstration present in
geometry? Even the geometry of antiquity, while being prior to any
"theory" of proof, saw a problem here. What characterizes geometry
is not its capacity to present its formations in sorne manner or
another; rather, this presentation must follow a completely deter-
mined process of establishment, using specific means of establish-
ment-it must take place constructively.
This is more than a mere technical problem. Properly understood,
this requirement is not concerned with precisely or correctly "imi-
tating" something already pre-given and known, using sorne techni-
cal means in a real medium; rather, the postulate of constructive
geometry contains nothing less than the geometrical formation, first
generated only in and through the construction. It only comes to
existence in this generation.
This state of affairs is ontologically important. Although in a
preliminary way, it contains the operationa1 sense of the being of
mathematical formation; its being is a being by virtue of a specific
method of generation. If the condition for its existence rests in the
possibility of its constructive generation, then the ultimate explica-
tion of its sense comes only from the constructive activity itself. The
illumination of this activity is an unavoidable task for any theory of
mathematical objectivity. Yet the illumination ought not to limit
itself to mathematical thinking, i.e., to the "constructive" aspect of
this thinking; rather, it must touch upon the fact that this latter
concept of construction is merely a late historical transposition of a
113. The assertion that axioms are undemonstrable principies is obvi-
ously more telling than that they are merely unproven presuppositions. The
former had fruitful results in the philosophy of mathematics. The concept of
the undemonstrability of euclidean axioms lent them a unique dignity,
initiating a plethora of speculations concerning the reasons why precisely
these and no other fundamental "facts" are undemonstrable. At the same
time, however, it was precisely their asserted undemonstrability that
brought geometry its progress, for when this assertion was placed in doubt,
it led to new attempts at demonstration; from the failure of these demon-
strations one did not gain an insight into the demonstrability of axioms, but
rather into the arbitrariness of an axiomatic system. Thus the undemonstra-
bility of the classical axioms ceased to be a metaphysical riddle; rather, it
acquired a relative significance. What is undemonstrable in one system is
not undemonstrable per se; rather, an axiom of one system can be a
demonstratable theorem in another.
Basic Trends of Mathematization
217
meaning having, even in mathematics, its direct roots in handcrafts.
A sufficiently extensive discussion of this suggestion would require
a more encompassing investigation in its own right. Here we can
touch upon sorne central points in order to suggest the specific
uniqueness of mathematical constructions.
Euclid's Elements already appear to be built up completely under
the aspect of constructability. For antiquity, this concept is totally
bound to specific means of construction-for example implements
such as compass and straightedge, which are to be employed in
accordance with established prescripts. However, the novelty of
Euclid's Elements does not rest, in the last analysis, on a strict and
systematically applied mastery of the two constructive implements
which were already employed, though unsystematically, prior to
Euclid. The constructive character of euclidean geometry shows up
primarily in that it employs only geometric formations whose
existence, while not explicitly postulated, is nonetheless demon-
strated by the possibility of construction with compass and straight-
edge.114 Whether the Greeks were clearly aware of the meaning of
their constructive process is obviously uncertain. In any case, this
constructive aspect will have to be touched upon if we are to raise
the question concerning the "essence" of the euclidean axioms. Why
did Euclid "choose"-if such an achronistic mode of phrasing is
appropriate here-these and no other axioms? A historically subse-
quent answer, in no wise supported by Euclid's Elements, is usually
offered with the notion that his axioms are "basic facts of intuition,"
comprising the "final intuitive evidence." Yet even with a concept of
intuition that can be explicated geometrically, this solution remains
insufficient, since not all euclidean axioms are adequate for such a
claim of evidence; sorne theorems derived from them are undoubt-
edly more intuitable than their axiomatic presuppositions. Euclid
himself does not offer any "meta"-mathematical remarks. Seen from
the structure of his entire system, it can be justifiably assumed that
these axioms were not made into first principies of geometry because
Euclid regarded them as intuitively evident; rather, the criterion for
their choice was seen only in the achievement of a constructive
building up of geometry, and one that is as complete as possible.
The following aspect of the mathematical fertility of axioms first
attained its full consequences in non-euclidean geometry. In it the
axioms are principies not only in the sense of being factually prior,
114. H. G. Zeithen has convincingly established the constructive character
of geometric objectivities. In addition, see A.D. Steele.
218 Mathematical Space
although subsequent for us, but rather of being at the same time
proteron pros emas and proteron te physei from which the building
up of the new geometry proceeds in accordance with strict axiomatic
deduction. The question concerning evidence for the axioms is for it
a totally nonessential, if not senseless question. The establishment of
the axioms is not determined by whether they are intuitively
evident; rather, only what they yield and what can be accomplished
with them is of significance.
The standpoint of "producing," of "establishing" the geometric
states of affairs dominates geometry in a novel way and extends the
concept of construction. Without its original meaning, construction
now means not merely the generation of geometrical formations by
means of specific implements, but also their constitution in pure
thought operations through step by step construction on axiomatic
foundations. Besides the pictorial construction of a formation from
intuitable parts with the aid of a few mathematical implements to be
handled in a determinate and prescribed manner, there appears a
construction from thought elements, which are to be connected by
firmly prescribed rules. While no longer bound to the instruments of
elementary geometry, the new mode of construction requires a
specific kind of working tool: the sign. The more geometry is
pervaded by sign symbolism, in contrast to the old means of
construction, the more the purely rational concept of construction
attains a fundamental significance for geometric existence.
New possibilities for the generation of geometry are thereby
opened-which, however, also leads to new problems that could not
emerge at all in the old geometry.
The predominantly new, in which the signitive symbolism far
surpasses pictorial intuition, can capture not only the pictorial and
picturable in its signs, but also the purely categorial such as logical
operations. This allows not only a signitive representation of geo-
metrical formations, but also a complete proposition about these
formations. In contrast to analytical geometry, a new isomorphism of
a higher level comes into play: the coordination of a sign system to
a propositional system. It is in this possibility that a formal theory of
proof can first be grounded. From its level of reflection and with its
signitive means the structure of mathematical thought itself can be
precisely grasped-yet at the same time difficulties brought about by
this new possibility of construction come into view.
It must be recalled that for antiquity the elementary process of
construction, employing the available means of construction and
the simple rules of their use, must have made it completely
Basic Trends of Mathematization 219
self.evident that a mathematical formation could be constructively
generated in a finite number of steps-indeed, self-evident to such a
degree that this finitude of geometrical construction did not even
have to be emphasized or affirmed in any specific proposition. In
contrast, the conceptual expansion of construction toward concep-
tual operations presented fundamental difficulties. Thus in arder to
guard against paradoxes, foundational mathematical investigations
found it necessary to posit a requirement that all formations that can
be generated ought to be attained in a finite number of construction
steps in accordance with specific pre-given rules of operation; if it is
to be comprehended in the most strict sense, this concept of
constructability contains this additional postulate of finitude.
Indeed, as it turns out, the finite modes of proceeding do not
exhaust all constructive means, although the finite means are
completely constructive.
The meaning attained by the requirements of finitude for the
constructive building up of mathematics allows the appearance of
the algorithmic structure of mathematical thinking. In its original
sense, an algorithm is universal, univocal by prescribed procedure of
calculation through which a solution to a particular problem can be
reached in a finite number of steps. If the concept of the algorithm is
extended beyond the common procedure of calculating to include
operating with signs in general, it can serve as a characteristic of its
system of rules. Each individual rule would then present an algo-
rithm, and the entire system of rules would also be conceived as an
algorithm to the extent that the use of rules is established univocally
and completely. It is only with the possession of these algorithms
that the genuine problem of calculability, decidability, and generat-
ability become present. Since they lead into specific foundational
questions of pure mathematics, they cannot be pursued here any
further.
115
Nonetheless, it would be beneficia! to touch upon Husserl's
delimitation of the algorithm and the definite manifold. For Husserl,
a definite manifold is characterized by a limited number of basic
presuppositions (concepts and propositions) which determine com-
pletely and univocally all mathematical states of affairs of the
provinces concerned so that all the states of affairs can be deduced
in unbroken sequence from the presuppositions; hence what is
mathematical or nonmathematical always follows as a formal-logical
consequence or as a logical contradiction from the axiomatic pre-
115. For more details see the works of P. Finsler and H. Hermes.
220 Mathematical Space
suppositions. Husserl explicitly indicates a close affinity between
his concept of definiteness and the axiom of completeness by
Hilbert. A finite number of axioms completely and unambiguously
determines the totality of all the possible formations belonging to a
province in the manner characteristic of purely analytic necessity, so
that of essential necessity, nothing in the province remains open,
which should accordingly distinguish mathematics from all other
sciences and constitute for it the decisive criterion as an axiomatic-
deductive science,llB
Two years after the appearance of Husserl's Formal and Transcen-
dental Logic, foundational investigations demonstrated the untena-
bility of the Husserlian criterion. Since then only the noncontra-
dictoriness can be maintained, and only if no limitations are allowed
for the means of its proof.
117
Today Gidel's proof of incompleteness
(for arithmetic), showing that the arithmetic axioms yield no deci-
sion concerning all arithmetic propositions, is a more serious
objection to Husserl's conceptions. Moreover, the Husserlian con-
ception of univocity, and the way that he has understood it, can no
longer be maintained.11s
116. E. Husserl, Ideen 1, 72; see also FTL, 31. O. Becker (2), pp. 686-89,
has presented the remarkable suggestion that Plato already had a clear
insight into the object of mathematics as a "definite manifold."
117. The freedom from contradiction of the pure theory of number cannot
be proven; it remains-with Husserl-"open," if we limit ourselves to the
means of a given system, as was shown by K. Godel in 1931. In 1936 G.
Gentzen succeeded in demonstrating the absence of contradiction for a
number theory only with the aid of transfinite methods, by using Cantor's
ordinal numbers.
118. In 1934, Th. Skolem was able to show that not only a finite number
of axioms but also infinitely enumerable axioms are inadequate to "charac-
terize" the number series. G. Martin, who has critically discussed Husserl's
concept of the definite manifold (in accordance with Ideen I), uses, among
others, the results of Skolem as a refutation of Husserl's requirement for
finitude; at the same time, he extrapolates from Skolem's work the proof
"that the arithmetic of the natural numbers cannot be grounded in a finite
system of axioms" (p. 163; italics added). lt remains to distinguish between
the axioma tic grounding of a mathematical domain and its characterization.
That the series of natural numbers is not "characterizable" by infinitely
enumerable axioms means only that its axioms are not specific for the
natural numbers, i.e., they do not univocally ground merely the number
series. This means that it is not possible to define numbers implicitly by
their use, but rather that there are other objects equally characterizable by
these axioms, such as certain functions belonging to number theory. In
Basic Trends of Mathematization 221
All these results are related to pure number theory and to
arithmetic, but not to geometry. Yet the incompleteness of arithmetic
ea ipso includes geometry since the corresponding proofs for geom-
etry can be achieved as a whole with arithmetic models. Husserl's
characterization of mathematics as a definite manifold of proposi-
tions can no longer be maintained; and insofar as it is an objective
correlate of an algorithmic structure of thinking, the latter, in light of
this characterization, also requires corrections. Mathematical think-
ing per se can be characterized as algorithmic only in a less restricted
sense; the truth or falsity of a mathematical proposition can no
longer always be decided through an algorithm; rather, the decidable
problem in mathematics can always be resolved by way of algo-
rithms. The only characterization that can be maintained today for
mathematics is that an infinite number of theorems can be acquired
as statements whose truth is based on a finite (enumerable) number
of premises-a state of affairs that says less than the Husserlian
conception. Yet in itself this state of affairs is sufficiently notewor-
thy. Thus it deserves not to be passed over as self-evident, but must
be understood as a unique contingency of mathematical thinking.
6. Summary
The preliminary observations of this section saw their main task in
the exposition of sorne modalities of consciousness that play an
essential role in the total nexus of mathematical intentionality as well
as in the phenomenological clarification of the modes of givenness of
its correlative sphere of objectivity. The immense complexity of this
nexus on the act side was not completely mastered, and the presen-
tations concerning abstraction and mathematical ideation, pictorial-
symbolic and categorial intuition, signitive symbolization and for-
malization, offered little more than an indication of problems
previously not yet treated phenomenologically.
Despite offering or neglecting to offer exhaustive analyses of the
foregoing intentionalities, sorne consideration must be given them
on two grounds. On the one hand, it was essential to extricate the
contrast, a separate question must be posed as to whether the numbers can
be grounded by a finite number of axioms. Under certain conditions this
question can be answered affirmatively if one envisages the axiom of
induction as singular, as it is used singularly in linguistic understanding. If
one wishes effectively to refute Husserl's requirement for definiteness, one
must show that this axiom presents a genuine, undisputed chain of axioms.
222 Mathematical Space
complex of phenomena constitutive of geometry, at least to the
extent of avoiding the danger of diminishing the fundamental
problems concerning the access to the geometrical domain of objec-
tivity. To cover up these problems with accustomed terms such as
"empirical" or "apriori," "intuitive" or "non-intuitive," "abstract"
and "purely logical" is inappropriate as long as, first, the justifiable
sense of these concepts is not properly clarified for geometry and,
second, as long as their limitations are not outlined as precisely as
possible.
On the other hand, the investigations achieved were important for
us, since they shed sorne new light on the subject who, as a unity of
consciousness and corporeity, pursues geometry. This unity is not
only factually unavoidable for a scientifically engaged subject, but
moreover obtains its structural illumination from specific activities
of "pure" scientific consciousness itself. Although at first glance it
may seem that in its orientation toward an incorporeal world of
objectivity, this consciousness far surpasses the lived body and all
corporeal functions, leading to the notion of a "pure" consciousness,
the complex of acts investigated shows precisely that in the con-
struction of this consciousness, corporeal functions take an active
part. In addition, it is inadequate merely to secure a founding
moment for mathematizing acts and their objective correlates in the
sensibly given "in person," as can be easily shown for the case of
ideation and categorical intuition. It can be made clear in the
phenomenal complex of intuition that the role of corporeity is not
given its adequate due if it constitutes merely a lower "stratum," or
something similar, in the domain of geometric objectivities. If
corporeity is to be taken with the full diversity of its various
characteristics of activity, then it is apparent that the image of the
lived body as the sale "source" of geometry is only very condition-
ally appropriate. Indeed, this image does not relate to anything
decisive in sensible intuition. It is not only in pictorial-symbolic
intuition that geometry requires corporeal functioning for its com-
pletion. Rather, paradoxical as it may seem, even in those formations
of mathematics that can extricate themselves completely from this
type of intuition and operate exclusively with signs, corporeity and
its functions are brought toa new validity. Modern constructivisms
are bound to the establishment of sign-series, not only factually but
also as the condition for their possibility. Their operating with signs
is not merely accidental; rather, the full actuality of this turn
necessarily takes place on signs. Finally, as Hilbert has basically
recognized, these signs are not "merely" and "also" sensibly perceiv-
Basic Trends of Mathematization 223
able images; rather, their constitution "in person" in sensible
intuition is of the highest mathematical relevance. This all reveals
that what consciousness can achieve in this new domain of objec-
tivity is possible only with the continuous participation of corporeal
activities. The lived body is a transcendental condition of these
accomplishments and, indeed, not merely a founding moment that is
surpassed by higher-level mathematical activities of consciousness;
rather, owing to the implicative intertwining of corporeal functions
and intentionalities of consciousness, it is also a condition for the
possibility of the accomplishment and completion of those activities
as such.
Perhaps we should speak here more cautiously: instead of "the"
lived body, we should speak of "a" lived body. Is it not the case that
modern mathematical theory is completely independent of specific
corporeal structure and resultantly of the human? It is thinkable that
at best this theory requires sorne corporeity and its functional modes
of generation of, and manipulation with, signs, without presuppos-
ing the specificity of these functions as those of human corporeity. In
such case its relativity to the human subject would be merely factual,
"also" demonstrable on it, while the constitution of such theory
would be thinkable in principie in terms of differently organized
corporeal being. Corresponding questions will reappear in the se-
quel, specifically in the mathematically oriented investigations; we
shall treat them when they emerge.
Finally, what is relevant to the specifically mathematical structure
of thought has often been judged negatively. It has been maintained
that mathematics is merely tautological. Insofar as mathematics
merely reveals what was already contained in its presuppositions,
the theorems follow analytically and formal-logically from the
presuppositions. This position must answer the question asto why
in mathematical deductions the same fundamental presuppositions
do not always yield equivalent statements, but rather extremely
varied and contentually different propositions which, in the case of
geometry, are crucial for the description of real spatial processes.
Indeed, the uniqueness of the geometric process consists of its strict
formal inference, and yet it also proceeds factually. Should the
ground for this be found in space itself? Should space itself enable
the specific material content of geometry? From what has been said
above, this could only mean that the first fundamental presupposi-
tions must be based in it and that they in turn are no longer formal
propositions, as if they were merely logical and thus subsumable
under the contentless category of objectivity, anything whatever.
224 Mathematical Space
Rather, they revert back to material formations containing factual
propositions of a kind such that in arder to be grasped with respect
to their specific material content, they require a particular reflective
access.
No mathematical theory of foundations and no theory of proof can
constitute such an access. Our subsequent observations will provide
questions and solutions for the existing individual researches. These
constitute our presuppositions. We accept them as given in the same
way that we accepted the things of lived space in the previous part.
Yet in the latter, we did not see our task as a mere ontic exposition
of the structures of lived spatiality; rather, we attempted to grasp it
ontologically in terms of the spatial understanding of the subject in
it and comporting himself toward it. Even with mathematical space
we are not concerned with a descriptive summary of what has been
worked out in specific researches. The following presentation does
not claim to be such a summary. lt can include only those states of
affairs that are relevant for the ontological consideration of geometry,
i.e., that aid in deciding the question concerning the "sense" of their
formations, their space. Systematic reasons dictate the placing of
euclidean geometry at the beginning of the following analyses. lt is
not only historically first but, in addition, it is ontologically the most
originary form of geometry, in a sense to be clarified below. The
point of departure of the following investigation may be compared to
that of the beginning of the first part of this work, as long as the
comparison is seen as nothing more than an external analogy: if the
investigation of actual space were to be a work that was appropriate
to the phenomena. lt could not dispense with determinations
belonging, according to their genuine sense, to subsequent forms of
lived space-after all, what was constantly given to us befare any
beginning was a subject who had already traversed all the spaces.
Thus here too, if the analysis of euclidean space is to be adequate to
the task, it cannot remain completely free from subsequent geomet-
rical determinations. Here, too, the subject who raises the question of
mathematical space as a whole, and begins with euclidean geometry,
is the subject who has already "measured" the totality of geometrical
spatialities.
SECTION TWO
Euclidean Space
Chapter One
Phenomeno1ogica1 Access to Metrics
l. Forrnation and Relationship. The Prirnacy of Relationships
The previous section contained various hints at sorne elernentary
geornetric forrnations which, while chosen arbitrarily, functioned
rnerely as exernplary rneans for the dernonstration of sorne essential
trends of the act-characteristics required in rnathernatics. At present
our concern is to consider sorne of these forrnations in their own
right. They and the relationships obtained between thern constitute
for us the elernentary dornain of objectivity of geornetry to which the
subject is oriented, just as the subject was oriented to the objectivity
of lived space.
Obviously this analogy should not be given too rnuch weight. In
lived space, the differing rnode of corporeal cornportrnent appeared
to be constitutive for the things of lived space in their differing rnode
of being; here the correlative relationship between the subject and
the new object-world seerns to be frozen into the one intentional
tension of the rnathernatizing consciousness. It has abandoned all
the lived connections to the world and has retained the one ideal
geornetric dornain of objectivity as its unique counterpole.
If talk about rnathernatical "spaces" is to acquire an acceptable
rneaning at all, then the justification of this plural usage cannot be
sought in the various orientations of the subject but exclusively in
the structure of such spaces and the objectivities that constitute
225
226 Mathematical Space
them, or in their various fulfillments of geometric meaning. This
objectivity thereby assumes particular importance. At the same time,
the following question must be addressed: How are we to grasp the
all-pervasive and constant mathematical intentionality of the sub-
ject in relationship to the multitude of mathematical spatial forms?
The question is not only what separates and binds them ontically-
to demonstrate this would be a matter of mathematics-but also
what constitutes their multiplicity ontologically, as seen from the
vantage point of the subject of mathematical space-conception.
The basic objects of geometry, in the sense of euclidean axiomatic
theory, are point, line, and plane, as well as the relationships
obtaining between them such as ... lies on ... (a point on a line),
intersection (of lines), and sorne relationship of arder (between, etc.).
The relationships can be called simple because their meaning is not
primarily and exclusively mathematical. They are understood and
sensefully diplayed in lived space; their geometric meaning is,
mareo ver, a carrying over of predetermined meanings into the ideal
sphere.
119
lt is otherwise with the aforesaid relata. Indeed, in lived space the
concepts of point, line, and plane have an intuitively fulfilled
meaning. Yet their genuine geometric content is not at all perceiv-
able in sensory intuition; rather, it first results from intuition
through a process of formalization. Here the concern is no longer
with a mere carrying over of the originally meant into another
domain. Rather, the intent is first and foremost to posit new objects.
What is meant with such objects intuitively, i.e., taken from the
domain of sensory intuition, is retained as a founding moment,
although with a decisive novelty which, in its constitution of the
"geometric," requires extensive discussion.
It is not by chance that the first attempts to ground geometry aimed
at determining the fundamental geometric formations through pre-
cise definition while the relationships between them, in contrast,
119. In the following, we differentiate between carrying over (ber-
tragung) and transposition (Transposition). The first directly accepts a
meaning given and grasped extra-scientifically into the geometric domain,
such that the meaning is made more precise and is given a strict delimita-
tion; in transposition ("translation") of a concept into the mathematical
sphere, there occurs a modification of meaning, yet in such a way that the
modification does not encompass the entire complex of meaning, while the
remaider maintain their unmodified meaning. (Cf. the mathematical concept
of movement, pp. 231 ff.).
Phenomenological Access to Metrics 227
appeared not to require their own definitive ascertainment. The
latter appeared to be obvious in their own right; lacking any specific
mathematical content of their own, they could be applied to geom-
etry without any special delimitation. Perhaps this is the reason why
Euclid lacks the axioms in which simple spatial relationships (such
as relationships of ordering) become thematic.
For a long time serious objections have been leveled against
Euclid's axiomatic theory. The insight into not only the practica!
uselessness, but above all the logical impossibility of explicitly
defining point, line, and plane was an important discovery of
foundational investigation. This led to the implicitly understood
definitions of these formations, which are possible only with the aid
of a preestablished system of axioms in which these formations are
not explicitly used. Hilbert demands that such an axiomatic system
should only establish formal properties of the relationships that
obtain between geometrical elements without explicitly stating what
the latter are.
The crucial difference between Euclid and Hilbert lies in that at
the very outset, Hilbert wishes to conceive of geometry as essentially
a special case of the pure theory of relations, with the relation
receiving primacy over its individual members. Yet in a very specific
regard, Hilbert's foundations of geometry do not go beyond Euclid's.
What is present in Hilbert's axioms of connection are the fundamen-
tal propositions of geometry used in the form of existential state-
ments ("there is ... ") of such relationships as "to lie on," "to go
through," "to join," "to intersect," etc., without prior definitive
delimitation of these relationships. The same can be said of the
second group of axioms pertaining to the relationships of ordering
whose description, as Hilbert says, "is best served by the term
'between.' " As already suggested, we are not concerned with a lack
of mathematical foundation, but with one that is found in the
essence of such relationships themselves. Their signification-
fulfillment is present prior to any mathematical sense-bestowing and
is taken over directly into the domain of mathematics.
120
The subsequent analysis of signification must be directed to
objects in a narrower sense, to geometric formations as such. We are
120. Hilbert (4), 2 and 3. This does not contradict what Hilbert says
about the axioms of ordering: they define the concept of "between." This
definition is implicit in the linear axioms of ordering, in which, however, a
pre-mathematical understanding of the verbal expression "between" is
already presupposed.
228 Mathematical Space
not hindered even if no explicit mathematical definition can be
devised for them. Our sale concern is systematically to trace the
various levels of meaning-giving in phenomenological description
with regard to the appropriate objects of sensory intuition.
2. The Line Segmentas a Fundamental Metric Formation
Seen from the vantage point of our concerns, point, line, and
plane, in the sense of fundamental geometric formations, are already
secondary and derivative phenomena. It is not only that mathemat-
ical meaning-giving, along with all the characteristics of activity
constitutive of it, reveals a higher-level intentionality in which the
directly grasping intentions are already presupposed; even the latter
are always oriented toward finite formations of sensory intuition,
whose boundaries cannot be arbitrarily transgressed either out-
wardly or inwardly. No points, lines, or planes are intuited here-
rather, things with "edges," limited by straight lines and plane
surfaces.
The edge is a morphological characteristic of the intuited thing.
Grasped as such and distinguished from other forms of physical
contours encountered on things by the touching and seeing corpore-
ity, it constitutes the sensible foundation for the conception of a
"line segment." The line segment must be se en as the genuine
fundamental formation of geometry, lending to geometry its origin-
ary sense as a science of spatial measure. The conception of the line
segment is prior to those of line and point not only historically and
factually, but also in its ontological meaning.
The conditions for its ideation are present in the sensory intuition
of the individual thing. As already mentioned with respect to the
space of intuition, its conception also includes the apperception of
the free motility of the intuited spatial thing and its invariance in the
transformation of its visual aspects. These are precisely the charac-
teristics of a spatial object determining itas a "fixed body," and they
are specifically discussed in various ways by modern physics. Here
we are not about to enter the debate of the individual sciences and
the easily misunderstood form in which they pose the question as to
whether there are "really" fixed and invariant bodies in movement.
Our concern is not with the problem of a fixed body envisaged from
a specific methodological viewpoint, but directly with its indubita-
ble phenomenon in the natural consciousness of objects, which is to
be taken and comprehended as such befare any special scientific
questions can be understood. The latter are possible on the basis of
Phenomeno1ogica1 Access to Metrics 229
a previously explicated concept of spatial bodily movement and
invariance. As has been shown, this is based on a possible world-
attitude of a sensibly intuiting and self-moving corporeal subject
who apperceives the thing as moving and, in this movement, as
changeless by virtue of bis own consciousness of identity and bis
own corporeal possibilities of movement. The fixed body, which
"itself" does not change in being carried and which is thus suitable
to be repeatedly "moved" for the purpose of measurement, is only
insofar as there is a motile and self-moving corporeity that knows
itself as self-moving and first experiences itself as "itself" in con-
fronting the intuited world of objects.
As a founding moment the corporeal movement in no wise
constitutes the entire nexus of conditions that are essentially re-
quired for the concept of a line segment. The phenomenon of the edge,
as boundary of a fixed body, is by itself insufficient to elicit the
specific meaning of the line segment. After all, what constitutively
belongs to its conceptual sense is its character of being a standard.
The line segment is determinable in principie in comparison with, in
its relationship to, another line segment. This means that only line
segments function as relata of such relationships. This reveals, on
the one hand, their ideal independence from the real world of things,
and on the other, their quantitative-relational character. Such origin-
ary, quasi-quantitative relations as "larger" or "smaller than ... "and
"exactly as large as ... " first assume exact expression and precisely
comparable relationships with the help of numbers. From these then
we obtain that one line segment is "contained in" another once,
twice .... Placed in such relationships, the line segment itself
functions as a unit of line segments. Once chosen and subsequently
taken to be the same and constant in all comparisons with other line
segments, it is capable of determining the sought for relationship of
size as a "relationship to itself," i.e., as a measure through repeated
application.
The ideational independence of a line segment, its extrication as a
pure and empty quantity from the world of material fullness, its
relational character wherein it attains its primary universal func-
tion in the domain of all that is measurable---.:all these cannot be
derived from corporeal functions; yet what the line segment is,
and what it accomplishes, can be understood only with these
functions.
The independence of the line segment does not appear only in
that, bound to the conception of a fixed body, it is the means for
comparing sizes within the sensibly intuited world of things. Once
230 Mathematical Space
conceived and grasped as independent, it pervades at the same time
the domain of free, ideal formations. What constitutes its authentic
metric aspect is not its pictorial-symbolically apprehensible form,
but rather that in it which presents the measurable and the measur-
ably comparable. Thus, for example, "side," "altitude," "diagonal,"
"radius," "axis," etc., are indeed concepts having a precise geometric
applicability (constituting, for example, a so-called position defini-
tion of a geometrical formation). Nonetheless, in their meaning
content they also bear morphological moments of form that cannot
be readily exchanged. They must be understood in relationship to
the whole of the formation. A square has no radius, and to speak of
a diagonal in an ellipse makes no sense. These enumerated aspects
can be mastered metrically only when their specific meaning within
the framework of the total figure is completely bracketed and the
meaning reduced merely to the conceptual sense of the purely
extensive size of a part of a line of a determnate length. It is only in
this respect that questions concerning the relationship between, for
example, side and altitude in a triangle, between the radius of a
circle and the side of an inscribed square, can become at all senseful.
Such relationships express the genuinely geometric content of such
formations; the relationships are given only between line segments,
i.e., between what is conceptually and significationally equal. This
leads to pure relationships of numbers.
The fundamental meaning of the line segment is not confined to
the objects of the elementary geometry. Even higher geometric
formations, such as algebraic and transcendent curves of higher
arder, become accessible in their characteristic metric properties
insofar as their relationships are measured against a determnate and
precise line segment. lt is irrelevant whether the line segment
remains invariant or is varied in accordance with sorne rule.
Even the determination of an angle is bound to the measure of a
line segment. While trigonometric functions and their inverse func-
tions contain transcendent numbers of relations, they do not change
the fundamental meaning of the line segment for the determination
of angles. After all, the problem of the incommensurable quantity
first attains its full mathematical import only in light of it. In
addition, the line segment retains its fundamental meaning in all
geometrical calculations of length and area. Tasks of squaring, of the
rectifiability of curves, have emerged, in their full extent, from a
geometric method that attempts to approximate the formations in
question by those of a more elementary kind. Thus the integral
calculus found its crucial problem in an appropriate formation of
Phenomenological Access to Metrics 231
limits which, in each singular case, had to proceed from the
elementary-geometrical, i.e., straight-line, through partial areas of a
finite size, bounded by line segments.
This founded meaning of line segment, assumed universally by
geometry, sheds light on the more recently devloped viewpoints in
mathematical investigations. The latter permit the expression of the
essential content of a geometric domain even in a single state of
affairs. What determines this content in its crucial outlines is the
relationship of line segments in contrast to determined mathemeti-
cal operations.121
But what does a "relationship" of line segments mean? Is it at all
thinkable that one line segment can relate otherwise than as one and
the same, wherever it may be measured? And what specific sense
can be given here to a mathematical operation that would allow the
application of this concept to ideal formations?
3. The Line Segmentas an Invariant of "Movements"
The first question involves the problem of a multiplicity of
geometries from which the geometry of invariant line segments, as a
special euclidean case, will be extricated; in contrast to the first, the
second question touches materially upon the method of geometric
investigation.
The inceptions for the latter are already found in Euclid, even if its
formal expression in mathematical group theory is of recent date.
The meaning of the line segment is best illuminated by a theoretical
aspect of geometry that is a fundamental component of geometry
both historically and systematically: the theory of congruence. We
are not concerned with the contents of the individual results of this
theory, but with the existence of congruent formations as such.
Congruence presents a more definitive relationship of equivalence;
as an equivalence of "coincidence" it means not merely the agree-
ment of two or more geometric formations in terms of spatial content
or form, but the agreement of both determinations at the same time.
However, the term not only concerns a specific geometric state of
affairs, but also expresses something essential with respect to how
this equivalence is to be thought and where the criterion should
ultimately be sought for its fulfillment in each individual case.
121. We are still avoiding speaking here of a line segment as a distance
between two points, since the concept of the mathematical point is not yet
clarified. See pp. 296 ff.
232 Mathematical Space
Moreover, the bringing to coincidence of the formations in this
question involves an operation that at first sight seems completely
unproblematic; yet it not only potentially includes a mathematical
prejudice of a very specific kind, but also comprises a significant
point of departure for the unfolding of the problem of mathematical
space.
This prejudice consists of the assumption that during the "pro-
cess" of coincidence, the formations will remain constant in their
form and size, regardless of the "way" in which the coincidence may
be attained. The metric formation thus exists independently of its
position. In other words, congruence geometry possesses formations
that are identical except for their location. This is not immediately
self-evident and does not touch upon a single possibility of a
geometric relationship. Hence the task is to obtain an insight into the
conditions for such a relationship of congruence. What are the
specific conceptual characteristics of the mathematically thinking
subject that would shed light on congruence? How can the geometric
sense of "congruent" formations be determined?
These questions revert back to our previous discussions of line
segment, insofar as the congruence of metric formations is deter-
mined by the invariability of its standards. But the line segment
appeared to be sensibly founded on the phenomenon of a fixed body
in the space of intuition. Freely movable and invariant as to its
length, this fixed body turned out to be relative in its being to the
subject of measuring and his own corporeal dynamics. It is signifi-
cant that this moment of corporeal motility is constitutive not only
for the apprehension of the "line segment" as a free geometric
formation, but also for mathematical operations. Although within
the subject matter of geometry "movement" has been explicitly
considered only since the development of group theory, it has been
seen as worthy of thought from its very scientific origins. This
conception is already found in Euclid's theory of congruence,
though implicitly, Euclid identifies superposable figures as such
through "movements."
122
It could be objected that the concept of
movement in geometry is only a figurative mode of expression or
122. Euclid, Elements 1, propositions 4 and 8. We shall not enter in any
detail into the controversy that flared concerning Euclid's treatment of
congruence propositions and his method of epharmasein. The frequently
discussed question as to whether with this method Euclid incorporated an
empirical moment into geometry by having recourse to "movement" is
irrelevant for the present context.
Phenomeno1ogica1 Access to Metrics 233
only a kinematic "form" of apprehending, used by a geometrically
engaged subject as an access into the geometric domain. Neverthe-
less, "in itself," as an ideal mode of being, this domain lacks any
kinematic and hence temporally-laden moment, and therefore must
be apprehended in sorne other more appropriate way. The concep-
tion of movement in geometry would be a scientific precursor, a kind
of heuristic principie without, however, corresponding to the es-
sence of geometric objectivity itself.
Meanwhile, our previous analyses have sufficiently shown that
any concept of being in "itself" can only be meaningfully explicated
through the incorporation of the immanent reflexivity of the con-
sciousness that apprehends being. What the geometric entity "may"
be "outside" this consciousness appears to be a futile question.
However, that this consciousness conceives of ideal entities in this
and no other way, as is shown through reflective analyses of
consciousness of ideal objects, is not only established as an ontic
state of affairs, but must also become ontologically understandable
in terms of the conditions of its possibility in the subject constituting
its being.
But this leads to the conception of movement in geometry and its
noteworthy interrelationships. First of all, it is illuminating that the
concept of movement be modified in its meaning if it is to be
applicable to entities that, in their ideal mode of being, can only be
thought as being outside of real interrelationships and the flow of
time. At the same time, there is no doubt that despite the modified
meaning, specific constitutive characteristics of movement as a real
spatial process must reappear in the geometric domain if this
concept is to be justifiable in geometry. First of all, the affirmative
aspects of this concept must be established. At the same time, this
will constitute an overview of the types of geometric movement as
su ch.
Things and their movement in the space of intuition appeared to
possess three characteristic determinations: (1) the identity of the
intuited thing in its changes of place, i.e., its independence of place
(place as a mere "location"); (2) the attainability in principie of any
place by paths chosen at will ("free" motility); (3) the feasability of
various types of movement. The first characteristic is equivalent in
meaning with the existence of fixed bodies in correlation to the
motility of the lived body as a "self." The free motility of the intuited
thing manifests the continuity and the homogeneity of space; in it
there are no "islands" that cannot be traversed and no places
distinguishable by their inaccessibility. After all, space is simulta-
234 Mathematical Space
neous for rnovernent and through rnovernent, narnely corporeal
rnovernent. It is the paradigrn of a continuous process as such,
which, in its capacity as a process, constitutes at the sarne time the
rnultitude of co-valent spatial locations.
Finally, the third property of rnovernent, to be able to be realized
in a rnultiplicity of forrns, can be understood only frorn the corporeal
cornportrnent of the subject. Just as each rnovernent of a thing is
relative to a self-rnoving being, and the forrner if characterized by
being cornpensated for in principie by self-rnotility, each moved
object can essentially acquire only those rnode types of rnovement of
which corporeity itself is also capable. The latter, however, can be
reduced to precisely two: rotating rnovement and linear movernent.
The first, in contrast to the second, requires a distinctive locus
conceived as being itself at rest, as a center of reference of rotating
movernent. If only such rotating movements were possible in the
space of intuition, then for each rnovernent in it there would have to
be a distinctive location and all paths would have to run back to this
location. Yet such a space could not be the space of intuition of an
intentionally structured corporeal subject. The latter is not rnerely
present sornehow as a unity of consciousness and corporeity; rather,
this unity is very specific and is given in a deterrnined manner.
While being precisely this lived body in its specific functional
organization, it is at the sarne time the lived body of an intentional
consciousness, which in its turn is co-determined from the outset, in
its structure and contents, by this lived body. Such a corporeal
subject would, however, find the continuous return to the same
place contradictory to the structure of its intentionality. In the
consciousness of a corporeity capable only of rnovement that returns
to itself, there would be, figuratively speaking, no "space" for its
possibilities of knowledge asan open horizon with a beyond for the
imrneasurable, the infinite, and the irrevocably unknown, with its
ever unknowable residuum. The difference between the known and
the unknown, between one's own and the alien, between the actual
and the "yet" possible, would merge. In their polarity, these oppo-
sites determine the tendency of progressing onward that manifests
itself in corporeity only by way of the translatory forward rnovernent.
Finally, the question must be raised asto whether sorne of these
properties of rnovernent are retained in the geometric domain of
objectivity. With regard to the first two characteristics of rnovernent,
the unchangability and the free movement of the object, both are eo
ipso ascertainable for the geometric domain. They are precisely the
characteristics that are constitutive for the line segment as a funda-
Phenomenological Access to Metrics
235
mental metric formation. With respect to the types of movement, the
space of intuition reveals the translations and the rotations. But
these ate precisely the types of movement that articulate geometry;
the group of movements (proper) includes the translations and
rotations and no others. The group property of geometric movements
has its intuitive foundation in the movement of real bodies. What in
group-theoretical calculus is called identity element ( or unit ele-
ment), inverse, and group product means here a state of rest
(identical mapping of itself), the cancellation of a movement through
its reversal which, in turn, is a movement; and finally the possibility
of executing several successive movements, whereby the end result
is again a movement.
What lends the concept of movement in the mathematical domain
its sense is the circumstance that specific characteristics of move-
ment that are determinant for the corporeally-centered space of
intuition are transposed into the geometrical domain. In other
words, with the complicated and multi-leveled enactment leading
from the sensory intuition of a real fixed body to the conception of
a line segment, these moments of apprehension remain related to
what can possibly happen with the fixed body and the ways that it
shows itself as fixed. Yet the conception of movement in the ideal
domain, as "transposed" from the real "movement," is in the
geometric area no longer a givenness "in person," nor is it an
immediate itself-givenness as it is in the physical domain. The line
segment does not move "itself," nor "will" it be "moved" by any
kind of cause as if it were a thing. But it is also not merely "thought"
as moved or, in analogy to the latter, merely "represented" as
movable. The latter may play a psychological role in elementary
geometric work, but it is not appropriate for the matter under
consideration. What transposition signifies here, and what it means
geometrically, becomes quite clear in the mathematical description
of such movements. It occurs in transformation equations that are
subordinated to specific formal conditions (e.g., the non-vanishing
determinants) whose common characteristic is their lack of temporal
variables. What these transformations express is, strictly speaking,
nothing other than a precise coordination of a formation with its
"mapping." It is a purely static correspondence of two formations.
Yet this does not simply repeat and trace over what was already
there; rather, it is of such a kind that with their appearance the
duality of the formations is first posited. Indeed, these equations
have the character of prescriptions whose fulfillment aims only ata
"mapping." There is no other way to think this except with the
236 Mathematical Space
conception that one formation has been "transmitted" into another,
has "moved" into it. In other words, what here motivates the
comprehension of movement is not a concretely grasped process
applied "on" the ideal formation itself-which would not be possi-
ble, as it is without temporal components; rather, it is a specific
mode of sense-bestowal of mathematically prescribed coordination.
This takes place through a being who is the subject of a lived body
with a specific corporeal structure of movement. That the extra-
temporal entity is based on a mode of apprehension of "pure"
consciousness, which nevertheless retains a moment of temporal
sense even in this ideal, extra-temporal domain, can be explicated
only through another specific objectivity consisting of the relation-
ship between a pure extra-temporal consciousness (i.e., conscious-
ness that constitutes something extra-temporal) and time-bound
corporeity. The domain of being constituted in the ideal "move-
ment" turns out to be nothing else than the objective correlate of a
particular type of transcendence of corporeal comportment. The .
kinematic mode of conception, as it has increasingly pervaded
mathematical science in various forms, can be understood only in
these terms. Such procedural thinking-whether it be called geo-
metric "mapping" or analytically designated as "transformation,"
under the rubrics of covariants, contravariants, and invariants-has
concepts that designate something more than would be merely
psychological, sorne mere process of thought; they designate some-
thing objective, signifying mathematical entities in their being.
4. The Concept of Movement as a Leading Concept of the
Theory of Invariants
The kinematic conception of geometric entities is most clearly
understandable in the general theory of invariants. Supported by the
calculus of groups, it is based on the crucial insight that the
numerous geometries cannot only be grasped in terms of a unitary
principie-as would be the case, for example, with the invariance of
certain geometric formations in contrast to specific movements-
formed from a subsequent extraction and compilation of aspects
common to them. Rather, at the very outset, each geometry is
completely determined through its invariants and thus through its
movements. The latter constitute the metric domain in the genuine
sense. In accordance with its structure, this domain is established as
soon as it is decided which geometrical quantities remain constant
in specific movements. Thus the invariance of the line segment
Phenomenological Access to Metrics 237
univocally characterizes the euclidean movements, and thereby the
sale "processes" in the geometric domain that justify the concept of
movement in its primordial signifigance. That a line segment would
change its length during a displacement contradicts the plain
conceptual sense of a displacement. To apprehend this event, the
conception has recourse to an additional moment that allows the
line segment to be "simultaneously" stretched or shrunk. Such a
process, which is the simplest geometrical example of an affine
(specifically of an equiformal) transformation, is, in accordance with
its meaning, no longer a movement or, rather, what is "moved" here
is no longer properly a line segment. In geometric thought, even this
process is constituted in no other way than in conceptual character-
istics whose ultimate founding level is located in corporeal modes of
comportment. In any case, it appears that these relationships con-
stitute a sense-bestowing core for all the higher forms of geometric
process, insofar as the meaning content of all of these forms is
constructed from the previously conceived meaning of a line seg-
ment and the sense of movement belonging to it, while at the same
time acquiring new sense. All ideal "processes" in the geometric
domain are significationally considered as variants of the primary
forms of process, namely the euclidean movements. But that is why
they no longer are geometric movements. In general they present
"mappings"; below them the euclidean movements constitute the
special group of congruent mappings.
What is significant is that each group of mappings, characterized
by the invariants that belong to it, completely determines a corre-
sponding geometry.
Just as the invariance of a line segment constitutes euclidean
geometry, so the relationships between line segments constitutes
affine geometry. Its transformation group is such that it transposes
line segments over line segments while generally varying their
lengths. Only the relationships among lengths remain, in other
words, what is moved in this geometry is not line segments, but
relationships among line segments. Yet this says nothing about the
angle enclosed by two line segments. If constancy is required for it,
in addition to the constant relationship of line segments, then we are
dealing with a specific subspecies of affine geometry, conformal
geometry. But since the angle can be changed variously by a
(universal) affine transformation, then morphologically speaking,
this geometry yields highly di verse formations proliferated by affine
mapping. In place of congruence there appears affine equivalence, so
well evidenced by the relationship between a circle andan ellipse.
238 Mathematical Space
While both formations are completely different in a metric-euclidean
sense ( one compares their definition through geometric determina-
tion of position with the help of the concept of distance!), they are
affinely indistinguishable. An "affine being" would know of no
formal differentiation of such formations unless it took the relation
of affine equivalence in the sense of our congruence. For us as
intuitive beings, this means that affine geometry impoverishes the
richness of euclidean-metric geometry by narrowing down the
manifold of forms. It subordinates itself under the law of its own
genesis, the group of affine mappings.
In their turn these are only special transformations of a general
"movement" group comprising the projective transformations. It is
characterized by the invariant of the dual relationship of four line
segments, which incorporates the affine invariant of the relationship
of two line segments as a special case. Here the concept of equiva-
lence attains a degree of universality transcending that of the affine.
The non-equivalence between the finite and the infinite that is
preserved in affine geometry is surpassed in projective geometry.
With the aid of the mappings in this group, it is possible to transmit
projectively the "infinitely distant" formation into the finite and to
make it geometrically available.
To go into more elaborate mathematical detail would be redundant
at this point. Our concern focused merely on the insight that the
various geometries alluded to here are justified in their multitude by
the necessity of the various invariants required. Their systematic
interconnection is, however, supported by the unifying methodolog-
ical viewpoint of the general transformational group. At the same
time they are determined ontically through certain kinematic be-
stowals of meaning, which on the whole are founded in the concep-
tion of mathematical "movement." The states of affairs of these
geometries in their totality contain indices pointing back to the
sense-constituting activities of a mathematical consciousness,
which even in its higher intentional achievements reveals its connec-
tion to and dependence upon corporeal functions.
Chapter Two
Euclidean Normal Space
1. The Concept of Mathematical Space
(Preliminary Conceptual Clarification)
The concept of mathematical space has frequently been discussed
in current times. The philosophical controversy concerning its
justification arase with the discovery of non-euclidean geometry.
But this controversy limited itself essentially to the use of spatial
concepts in these disciplines without coming to any genuine deci-
sion. Lotze's proposal to speak of spaceoids in non-euclidean
manifolds in contrast to the space of euclidean geometry could not
take root. It presented nothing more than a terminological compro-
mise; adequate clarification concerning what we are to understand
by space as such in geometry was still lacking. The acceptance as
well as the rejection of his proposal lacks to date any discussion
focused on the subject matter itself. Whoever denies that geometry is
a science of space, and wishes to see it deal only with formations in
space, must no less justify the concept of space he employs than the
one who unreflectively applies this concept in geometry, above all
when he cannot avoid the conception of a multitude of mathematical
spaces. Definite expositions and delimitations of this question can
be purposeful and, indeed, are unavoidable in the work of the
special sciences, yet they lead no further toward understanding what
is to be explicated here. A conceptual clarification basing itself on a
strict and material phenomenological analysis can and must avoid
any terminological debates for or against a mathematical space. Its
only appropriate question concerning space is completely neutral:
its sale interest lies in the fact that in geometry one speaks of space
and spaces. Accordingly its only task is to explicate what constitutes
the characteristic content of the meaning of such language. The
question is not whether "one" ought or ought not to concede
togeometry its own concept of space. For us such an anonymous
239
240 Mathematical Space
"one," such a subject, does not exist. There is only a science of
geornetry which claims for itself a concept of space. In turn,
however, it is a fact that even this claim is historically quite late, and
that "space" was notan object of investigation for geometry frorn its
historical beginnings. The appropriate question for us is therefore:
what was the rnotivating reason that led to conceiving geometric
objectivity not only as an independent dornain in its own ideal rnode
of being and its own structurallaws, but also as an objectivity of an
ideal space? Conceived in the broadest terrns this rneans, after all, a
way of speaking about geornetry as a science of space. For it, space
is not sirnply the extensive manifold of real things and their
relationships (regardless of how it rnay be "applicable"); rather, as a
free science, geornetry is concerned only with itself. It is not directed
toward possibilities of application. It constructs a concept of space
that-in accordance with its sense-is free frorn any meaning of
sorne real space of objects. Hence the concern here is not with any
question of applicability of geometry in lived space; rather, the
problem of mathematical space must now be posed only for free
geometric manifolds. It must thus be taken as an ideal-ontological
question.
The unquestioning and unhesitating way in which modern geo-
metrical thought carried along such a conception of space into
mathematical formations must not mislead us into thinking that we
are dealing with something obvious, i.e., with a conception of space
that is irnplicit from the outset in geometrical work. Rather, it has to
do with a positing that is not yet grounded in pure geometrical
objectivity as such. Thus even geometry, in its beginnings, did not
conceive of such a space. Although Euclid established geometry as
an exact science, a strictly geometrical concept of space was still
cornpletely foreign-the geometric "euclidean space" is not a cre-
ation of the founder of euclidean geometry.
Nonetheless, euclidean space is not justa complement to thought,
a conception that merely supplements that of geometric objectivity
and that could be completely absent; rather, it has throughout its
fundamentum in re. That modern geometric thinking sees its forma-
tions "embedded" in an ideal space, as we may provisionally
formulate it, is a phenomenon of mathematical consciousness that
not only cannot be contested-for as such it offers a critical point of
departure for the rejection of any conception that would wish to
discredit an ideal spatiality-but in addition signifies a factual
dependence of this space on geometric objectivity and its relation-
ships to it.
Euclidean Normal Space 241
From what has been said, it is easy to understand what constitutes
mathematical spatiality. If the geometric objectivity is presentas an
"event" in a process of executing determinate transformations, if the
metric proper to this domain is conceivable only on the basis of an
ideal possibility of movement of its formations, then objectivity
conceived in this mode obviously requires an ideal "wherein" of
such movements, an ideal "where" of their occurrence. This is
posited, although not explicitly, when we speak of geometric move-
ment, or even when it is merely grasped as a depiction in the process
of thought. That there "is" an ideal space is equivalent to the
statement that the geometric entity reveals its fundamental metric
characteristics only in specific movements.
This explains a thought-provoking analogy obtained, on the one
hand, between corporeal movements and lived space and, on the
other, between geometric movement and ideal space. In the first
case, the filledness of space in its specific mode of being-and
thereby the corresponding space-is constituted in a strictly recip-
roca! correlation with a dynamic mode of comportment. In the
second case, the ideal domain, the correspondence to lived space is
found in that geometric space is not something befare its "content,"
but rather is first built and opened up by the geometric formations or,
to speak more exactly, by the given transformation group that
determines this content. Here, as well as in the former case, the
different movements are not merely processes in space, but also-
when one bans all genetic conception from one's language-"gener-
ators" of space.
In the case of geometric space, this concept of generation has a
most palpable meaning. lt is not concerned with finding a transfor-
mation group correlated to a somehow already conceived ideal space
that is meant "in itself," such that the group functions to describe the
relationship of the geometric formations "in" this ideal space.
Rather, it is solely the group, which is primary and established in
advance, that decides about the existence of an ideal space at all.
Each group of mappings projects a specific mathematical space; this
state of affairs comprises the meaning of all talk about euclidean,
affine, and projective space.
It is no wonder that within the special sciences, the existence of a
mathematical space first carne to conscious insight only with the
construction of analytic geometry. The Cartesian discovery that
every geometric formation can be related toa "straight line" signifies
the first explicit postulation of a coordinate system. Ever since, every
geometric proposition is related in a specific way to a system of
242 Mathemati9al Space
coordinates from which mathematical concepts such as "position"
and "location" obtain their precise geometric meaning.
The postulation of a coordinate system shows clearly what binds
the space characterized by such a system to lived space and what
sharply differentiates both types of spatiality. With respect to the
ontological meaning of such a system, there is an appertinent
objectivation of that which allots to each thing its there, its place in
relation to other places, within lived space: the corporeal here and
its directions of orientation. As a coordinate system [Bezugssystem]
for all topographical assertions in ideal space it is, to speak with
Weyl, "the ineradicable residuum of corporeal subjectivity" in a
space devoid of corporeity.
123
Even mathematical space is space only
for a consciousness and by virtue of a consciousness, which indeed
does transcend the space-determining and primordially space-
constituting functions of corporeity without setting them completely
out of play. While being objective, in the sense of being exclusively
an object for an intentional consciousness, even mathematical space
as "space" is unable to be constituted in any other way than through
determinations having their ultimate founding moments of meaning
in corporeity-and in the corporeity of a subject constituted in this
and in no other way.
Even the preference for right-angled coordinates and the experi-
ence of their simplicity must be founded here. The right angle
possesses, in contrast to any other measure of angles, a characteris-
tically exceptional position. In geometry this is clear from the fact
that the traditional division of angles presupposes the viewpoint of
the right angle, which is posited as a primitive standard, while the
rest of the angles are characterized as smaller or larger in comparison
to it. Seen phenomenologically, orthogonality presents itself as a
formalization of the absolute opposition of directions of movements,
as they differentiated themselves in the gradual succession of
corporeal dynamics, determining at the same time the structure of
the corporeal space of movement in accordance with surface and
depth.
Coordinate system and movement group are genuine conceptual
moments founding the meaning of mathematical space. That "there
is" mathematical space, that the conception of its existence is
meaningful, depends completely on both moments, which deter-
mine the mathematical concept of space and establish its limits.
Neither moment is independent from the other; rather, they have a
123. H. Weyl (1), p. 72.
Euclidean Normal Space
243
specific interconnection with each other. lt is only with the aid of a
coordinate system that the geometric formation becomes fixed with
precision and its movement becomes analytically describable. The
interconnection between coordinate system and geometric move-
ment had already been recognized by Descartes with regard to its
fundamental principie. lt could be objected that the Cartesian
"straight line" introduces into the science of geometry something
that is outside of and thus arbitrary, accidental to the formation.
Descartes counters this objection by pointing out that with the
continuous transformations of the coordinate axes (noted only by
him), the formations change their position, yet they themselves
remain "of the same species"-and for Descartes this appeared to be
their essential geometric aspecV
24
Even geometric movements are relative movements in contrast to
a coordinate system that is taken to be fixed; each movement of a
formation can be made retrogressive by a corresponding counter-
movement of the system. In a more recent mode of expression, the
invariants of a given transformation group are nothing else than
invariants in contrast to specific coordinate transformations. As such
they are primary for the genuine sense of the metric: in the more
precise sense of the "geometric," the only property that is valid, in
contrast to coordinate transformations, is the invariant.
However, what distinguishes mathematical space from lived space
is also most visible in the coordinate system and the mathematical
movements. That the coordinate system is freely choosable, or that
the geometric formations can move freely in contrast to the system,
is still a possibility presented as a transposition of the modes of
corporeal movement, although the motivation of movement is fun-
damentally changed. In lived space, corporeal movement takes
shape in forms and orientations in accordance with the different
sense-content of the lived body and its situation; the lived body
attains this sense-orientation primarily from something else, which
assumes a specific material determination and qualitative fullness
through its own mode of being. Yet in a space without corporeity, the
motive for movement is purely rational, and indeed rational in a
specific mode. The simplicity of a formation, as an aim for a
mathematical investigation in the choice of a coordinate system,
124. R. Descartes, Vol. VI (La Geometrie, Livre Second), pp. 388 ff. What
is crucial in this discovery ought not to be overlooked simply because
Descartes does not yet strictly distinguish between generalization and
formalization.
244 Mathematical Space
corresponds to a specific principie of economy that is purely
quantitative and in no wise geometric; it is determined by the rules
of the domain of numbers. The simplicity sought for is not related to
the formation as such, which is fixed in its geometrical form and its
metric characteristics, but rather to its analytic "form." By accepting
number as something completely distinct from the spatially struc-
tured domain of quantities, geometry has introduced viewpoints
completely foreign to space.
What is important here is that the analytically-algebraically con-
structed geometry enables a dimensional extension of mathematical
space. Previously the question of dimensions was completely over-
looked, since implicitly one assumed that its three-dimensionality
was similar to that of the natural space of objects. This three-
dimensionality is nevertheless not accorded any special status in the
framework of free geometric science. As a consequence of analytic
geometry-i.e., a geometry that no longer has its "forros" in spatial
imagery, but only in algebraic coordinations of signs that are
investigated in terms of pure algebraic attributes ( degree of equa-
tions, number of parts, etc.)-the question of dimensions becomes
meaningless for its concept of space. Once a coordinate system is
created, there is the possibility of symbolizing each point through
numbers and of formalizing the distance between two such points
(whose meaning in elementary geometry is that of a line segment) in
a quadratic expression, so that only the degree of this equation has
geometric relevance; there is then nothing that can hinder the
interpretation of the multi-membered quadratic equations geometri-
cally as "line segments." The "movements" of these "line segments"
can be meaningfully extended from the pictorially intuited domain
of two- and three-dimensionality to span over a correspondingly
dimensioned space, within which all the propositions of euclidean
geometry demonstrably hold.
The meaninglessness of the question of dimensions for geometry is
most striking in the sign symbolism of vector calculus. lt functions
without any fixed number of vectorial components, and its symbol-
ism is so construed that these components do not even appear in the
calculus. Only a subsequent limitation may freely interpret the
completed calculus as "meant" for S
3
, S
4
, Sn. For a vector-
analytically constructed geometry there is always the free possibility
of grasping itas the geometry of space possessing an arbitrary (finite)
number of dimensions. This has to do with the possibilities of
extension, which can easily be misunderstood as a generalizing
universalization. Yet the geometry of n-dimensional space is no
Euclidean Normal Space 245
more universal than that of three-dimensional space. Insofar as n
stands for any arbitrary natural number, it does not subsume
something under itself as a genus would subsume a species, but
rather contains it as a singular example of itself. A six-dimensional
space is no more universal than a three-dimensional space, just as in
the series of natural numbers a larger number is no more "universal"
than a smaller.1zs
Similarly, the ten-dimensional euclidean space is no more abstract
than the three-dimensional. What legitimates this common mode of
speaking is the everyday and vague signification of the abstract in
contrast to something that is intuitable pictorially or representation-
ally. A closer explication will show in what sense one must speak
here of intuitability or non-intuitability (p. 253). In any case, the
transition from a three-dimensional geometry to one which is
higher-dimensional has nothing to do with a process of abstraction.
If one observes strictly what is completely and genuinely contained
in the sense-bestowal of such mathematical spaces, and the mathe-
matical intentions in which they are constituted, then their concep-
tion as abstract spaces finds no support-just as one could say that
there is no abstraction when in an act of free choice we ascribe m or
n components to the generating vectors of a linear space. The mode
of speaking of euclidean "spaces" says nothing more and nothing
less. To reject this plurality as senseless would mean to reject even
the singular-and yet at the same time it would mean to deny the
existence of an analytic geometry as such.
An important supplement must be noted. Insofar as the previously
sought clarification of the concept of mathematical space has incor-
porated the diverse possibilities of dimensioning, it is supported by
a vry specific case of mathematical spatiality. What determines the
structure of mathematical space is not solely its geometry, but also
its topology. Thus even if the euclidean geometry discussed till now
has as such a univocal metric determination, no one specific
mathematical form of space has yet been decided upon univocally,
insofar as various possibilities . of connexus remain open. The
125. In a remarkable way Husserl seems to maintain the generic constitu-
tion of the various dimensional (euclidean) geometries. In accordance with
LU, Prolegomena, 70, S
3
should count as the "ultimate ideal unity" in the
genus of "purely categorically determined manifolds," which for Husserl
subsumes the geometry not only of euclidean, but also non-euclidean
spaces. Here, however, it would be appropriate to distinguish between
generalization and formalization.
246 Mathematical Space
simplest example is offered in the two-dimensional domain where
we compare an open plane with the surface of a cylinder. Both are
metrically equivalent, since the geometry of planes is valid for the
latter, yet they differ in their relationships of connectedness. Hence
the investigation of a mathematical space requires the discussion of
metric and topological relationships.
This state of affairs does not limit the space-opening function of
movements in geometry. While strictly speaking they are generators
only of metrics, they nevertheless play an equally significant role in
topological relationships since topology and metrics do not vary
independently of one another. In the special sciences, the more
detailed characterization of a mathematical space is taken almost
exclusively from metrics. Thus, for example, the surface of a
cylinder, despite its deviant connexus, is seen as euclidean space;
the same is maintained for the Klein-Clifford forms of space of which
we shall speak later. We now enter into this mathematical domain.
2. Normal Space (Euclidean Space of the Topological Type of
the Open Plane)
With the appearance of topological questions a new complex of
problems comes to the fore for our investigations. In light of this, the
investigation must face the task of understanding the appropriate
connexus which would be the only one posited originally in the
geometry determined through the euclidean movement group,
namely, the topology of open-endless space.
126
It is thus to be asked
whether in historical retrospect those topological possibilities that
euclidea:n metrics leave open in principie must be seen simply as
fortuitous facts-i.e., whether for the formation of the mathematical
concept of space, only those possibilities were originally selected
that were required for the type of the so-called open plane-ar
whether this "normal space" which, despite all subsequent con-
structions of other mathematical spaces makes no claim for geomet-
rically privileged position, can be conceived in terms of the subject's
activities-which precede all individual geometric work and lend
126. Following O. Becker (1), for the sake of brevity and terminological
univocity, we designate the euclidean space of topological type of the open
plane as normal space; It is only in terms of this terminological delimitation
that the concept of anomaly is subsequently to be understood in the sense of
deviation of other ideal spaces in a metric or topological respect (metric and
topological anomalies).
Euclidean Normal Space
247
intelligibility to topology, and from which the latter obtains all
sense-clarification.
The solution of this task requires a deeper inquiry into the subject
matter in arder to show that the elementary geometric formations are
founded in sensibly intuited factors. It is necessary, therefore, to go
back into this founding stratum and its specific characteristics.
Sensory intuition is depth-intuition-understandable ontologi-
cally as characteristic of an orientation toward the world that is
possible only for an intentionally directed corporeal being. In
addition, it is a finite intuition. As a mode of comportment of a
corporeal subject essentially characterized by finitude, it is limited
by a horizonal domain of space. It will become obvious that the
distinctive structural properties of normal space, linear and plane,
are accessible from here; they lead to the understanding of its
topology.
At the outset, the relationships here are different from the
relationship between an edge and a line segment. Line and plane
not only do not appear in the space of intuition, but there is not
even a simple and obvious relationship of correspondence to
sensibly apprehensible objectivities, as would be the case for an
edge or a line segment which are both characterized by their
limitations. Thus to understand a line we can neither have recourse
to a line segment nor think of it as emergent from arbitrarily
repeated line segments. Naive thinking assumes that it can take
refuge in the view that it is possible to extend the line segment in
both directions as often as it pleases, even while also denying the
actualization of such a possibility of extension. Meanwhile,
rectilinearity is already presupposed in the apprehension of a line
segment; on the other hand, the very assumption of a merely
potentially infinite progress from line segment to line is misleading,
insofar as such a process would result in an arbitrarily long line
segment, but never in a line. A line is not an arbitrarily long line
segment since the difference between a line segment and a line is
not a difference of measure. Regarded critically, the frequently
employed terminology that a line possesses "infinite length" states
nothing more than that here the concept of length has no
signification-fulfillment. The line does not "transcend" but rather
"subscends" all measures. It is not a metric but a topological
concept. Analogously, the same holds for the concept of the plane.
To think it as originating from an arbitrary addition of limited plane
parts would evoke similar reflections. Like the line, the plane is also
not a metric but a topological formation.
248 Mathematical Space
This reveals that we have retrogressive indices toward founda-
tional moments that are not to be located in sorne individually given
aspects of sensory intuition; rather, they must be sought in the total
structure of the space of intuition as such.
The space of intuition is structured in terms of extension in depth
and width. The depth of corporeally-centered space is formally
distinguished from extension in breadth, insofar as it alone pos-
sesses a univocity of orientation. While in the extensional spread
two phenomenal points can always be related in manifold ways and
modes, in the extension of depth this is possible only in precisely
one way. This univocity is founded in the particularity of depth
which is to be one behind the other. This being behind one another,
in contrast to being next to and above one another, acquires its arder
directly from the location of the intuiting corporeity. At the same
time, corporeity has a univoca} criterion for the fulfillment of this
arder in what phenomenally constitutes the "coincidence" of two
points lying one behind the other. Indeed, this coincidence is not an
intuitive, but a purely visual givenness. In contrast to the visual
field, the intuitive apprehending includes the apperception of the
being-one-behind-the-other, despite the visual coincidence of the
two points. But this is the constitution of the primal meaning of
rectilnearty. The intuitive glance would encounter any series of
phenomenal points in the characterized position "straightway" if it
could penetrate the ones in front.
How significant is the particularity of such univoca! relationships
of location, how dominant is the depth dimension in its structural
preeminence for the total apprehension of vision, is revealed by the
concept of the visual "ray", whose meaning emerges from the depth
dimension of the intuited world. It must be kept in mind that with the
ray, the domain of the phenomenally given is already abandoned. The
intuiting corporeity does not have a bundle of rays befare itself, but
a spatially structured world of things with a dimension of depth. It
does not see rays but things in space. The ray-character of vision is
not itself given for vision and, thus, phenomenologically speaking, is
nota characterization of its functional structure as given in corporeal
comportment; rather, it first acquires its meaning through an objec-
tivating inspection of the perceptual event. However, that the visual
function is the only one among the sense functions that can be
objectivated in this manner, that it can be fulfilled in the objectivating
apprehension as a "ray-like" event, is not at all an empty construction
of experimental optical science, but rather has the conditions for its
possibility in the structure of the sensibly intuited world of objects
Euclidean Normal Space 249
itself, as a world dimensioned in depth, and at the same time in the
specific functional organization of the corporeal subject.
Apart from the depth dimension, the analysis of the space of
intuition has shown the other essential structural moment of this
space: extension, with its degree of manifoldness one step higher
than that of the depth dimension. Opposed to the univocal and
corporeally-related intentional depth, extension appeared as a man-
ifold of pure relationship of things. Together with depth, it consti-
tutes the structure of a space in which the intuiting corporeal subject
exists in an ambivalent state: on the one hand, as a being "in" a
situation, and on the other, as a being "over against" a constellation
of states of affairs. As such, the pure two-dimensional extension is
no less relevant for the entirety of the sensibly intuited space than
the one-dimensional extension into depth.
Extension, to be sure, does not necessarily mean a plane exten-
sion. The ambiguity of being next to one another (in complete
contrast to the univocity of being one behind the other)-more
precisely, the possibility of relating two closely located phenomenal
points in various ways-does not motivate the apprehension of
planarity, but rather constitutes the universal meaning of two-
dimensionality. In contrast to it, the plane is a very special case. How
are we to conceive the unique meaning of a plane surface as well as
of rectilinearity in contrast to curvature? How is it that in our
intuition and conception the plane functions as a kind of fundamen-
tal surface and that in contrast to it, any other kind of curved surface
is grasped merely as a "deviation from" the plane?
It is insufficient to point to the structural simplicity of the plane in
contrast to the curved surface; this merely shifts but does not salve
the problem. Still to be clarified is how the concept of simplicity,
regardless of how it is understood, is related to a surface with plane
determination.
This state of affairs is comprehensible only with the aid of the
structural determination of the visual function. The apprehension of
planarity emerges in a process of objectivation that allows vision to
be conceived as a ray-like event: the "wandering" of the "gaze," as
a mode of comportment of a corporeal subject, is reduced to a visual
ray that "spreads over" a surface. But since this visual ray is formally
generative of the surface, it also determines the degree of curvature,
and thus the surface must necessarily be a plane.
127
In the case of a
127. We are not overlooking the fact that we are dealing strictly phenom-
enologically with determinations of a specific realm in terms of a "wander-
250 Mathematical Space
gaze ranging over a curved surface, the visual rays would have
merely a tangenitial function and, thus, roughly speaking, the whole
would fall outside of their range; this would not be the surface that
would be generated by the wandering of the gaze.tza
In addition, the concept of similarity of the forms of things in the
space of intuition thereby attains an intuitively fulfilled sense. The
existence of similar things, i.e., similar in form although varied in
size, is essentially bound to a plane-structured spatiality, since the
constancy of the angles of the formation, despite the changes in
extension, are possible only in it. It is possible to conceive of a
corporeal being whose functions are of such a type that no univocal
depth and plane extension would be present in direct intuition. For
such a being, no forms of things would be similar in its space of
intuition. In contrast, for a subject of our corporeal constitution, the
plane functions as the bearer of a basic topological determination of
everything extended.t
29
(As an originary spatial apperception,
founded in the functional determination of the lived body, it has a
dominant structural movement of representation beyond what is sen-
sibly intuitable, where thought turns to spatial metaphors. Thus the
partners in a discussion encounter each other on the same "plane" of
thought although they do not find themselves on the same "surface"
of thought. And should their conversation remain merely "superfi-
cial," the representation of the plane, in this phrase, also plays a role.
ing gaze." Of course, in a preliminary way this also touches upon
rectilinearity and planarity.
128. The theoretically conceivable case of ruled surface, i.e., of a surface
that has lines in specific directions, yet in general has a degree of curvature
other than zero, is here excluded, since the wandering of the glance is not
subtended by any limitation of direction and can be enacted arbitrarily.
129. The interconnection of planarity and similarity has been shown
repeatedly in mathematics. Bolzano has employed it while demonstrating
the plane-structuration of space in terms of the phenomenon of similarity.
The mathematical equivalence of both concepts should not obscure the
circumstance that ontologically speaking, we are concerned not with a
mutual conditioning, but with a one-sided relationship of dependence; the
meaning of planarity or of rectilinearity is prior and cannot be derived from
the concept of similarity. The space of intuition "must be" plane structured
not because in intuition there are similar things; rather, it can contain
similar forms because it is structured by a corporeal being of the given
functions. Apprehensions of similarity in the space of intuition ha ve already
presupposed the apperception of the line, because without it, the
signification-moment of constancy of angles, which is constitutive for
similarity, could not be intuitively apprehended.
Euclidean Normal Space
251
While rectilinearity and planarity, as spatial structural moments,
are completely based in the sensibly intuited world and its
corporeally-related arder, the notions of line and plane themselves do
not yet come into view. However, in light of our point of departure,
both are easily understandable in their own conceptual sense. We
recall that the mode of givenness of the space of intuition is deter-
mined from the start by the co-presence of "the" one space, so that
the former is originally apperceived as a "mere section" of the latter.
Yet this space is homogeneous and open-endless; it is conceivable
as a univocal continuation of the thing-relationships predominating
in the finite domain. The characteristic relationship among things
contains the experience of being one behind the other in spatial
depth. In accordance with this idea, it can be continued endlessly
and for the sake of univocity it is posited as an open-endless
progression. At first, this progression assures only the infinity in
principie of the visual ray. Through the homogenization of this
space, the center, as a point of inception of the visual ray, is reduced
to any phenomenal point of space and no longer has any special
significance. A point among other points, not only can it be arbi-
trarily shifted, but also, in a strict conception of the homogeneity of
space, it can and must completely forfeit its special character as a
point of departure for something. At the same time, the "visual" ray
becomes redundant; what remains is the purely formal meaning of
rectilinearity. It no longer includes the co-positing of a corporeal
function and, while being without any point of inception, it is
conceived as open-endless in two directions, i.e., as a line. In its free
motility it is generative of the all-sided and open-endless plane as a
structural element of the homogenous space of objects. That it is
open-endless means nothing other than that it is a space structured
in terms of planes.
The fundamental meanings of line and plane condition our con-
ception of a fundamental formation of euclidean geometry or
euclidean metrics-the line segment-in terms of rectilinearity. This
must not be taken as if the apperception "line" and "line segment"
were to occur in separate acts of apprehension. Rather, the conceptual
sense of the line segment originarily contains a moment of rectilinear-
ity justas the marking off of line segments in the process of measuring
is thought of as a continuous rectilinear process. This conception is
completely independent of the factual operation of marking off which
in individual cases may deviate considerably from the straight line.
What is important is that in such a marking off, the rectilinearity of
the measured path is constantly anticipated, and any deviation from
252 Mathematical Space
it is seen privatively, to be determined entirely in contrast to
rectilinearity.
This means, however, that euclidean normal space is a plane-
structured space. In accordance with its mode of generation through
the euclidean movement group, with its characteristic invariance of
the line segment, it guarantees e o ipso the free motility of a formation
structured by rectilinearity, leading from the very beginning to the
conception of an open-endless space.
The question raised at the outset of these paragraphs concerning
the special phenomenological position of euclidean normal space
among other mathematical spaces is thereby answered. The answer
consists in nothing other than in its topological congruence with the
space of the natural consciousness of objects.
This congruence is explainable in terms of the simple relationship
of founding between its constitutive formation, the line segment, and
the intuitive form-characteristics of the real-spatial world of things.
In light of subsequent distinctions, it is essential to introduce here the
term of immediate founding. This immediate founding of euclidean
normal space in the natural space of objects establishes for euclidean
space-both metrically and topologically-a certain preeminence.
Insofar as the latter must be seen as founding for any kind of space
in mathematics, "euclidean space," i.e., the open-endless manifold
constituted through the euclidean movement group, should be con-
sidered as the most primordial of all mathematical spatialities. As was
shown above, this primordiality is distinguished not only in its his-
torical and factual priority, but also in its ontological determination
of the very sense of existence of this space. It is grounded in the
structure of the sensibly intuited spatial world, and obtains its ulti-
mate foundation in the specific nature of the corporeally bound and
corporeally conditioned achievements of consciousness of a subject
capable of thematically meaning space as an object.l
30
3. The Question of Intuitability in Euclidean Geometry
The interrelationships discussed in the previous section offer the
possibility for clarifying the conceptual sense of intuition in geom-
130. The particularity of the euclidean spatial structure is, to stress once
again, not mathematical but ontological. In the constitution of the multitude
of mathematical spaces, euclidean normal space functions as a kind of
primordial objectivity from which subsequent spaces first obtain their
meaning as "spaces." See pp. 297 ff.
Euclidean Normal Space
253
etry. lntuitability in euclidean geometry now more clearly means
intuitability in the geometry of the type of the open plane. The
intuitability meant here is hereafter appropriate for both metric as
well as topological relationships.
Of course it is tacitly assumed as obvious that the concern here is
with a kind of intuitability that is distinct from simple sensory
intuition. We can point to our earlier expositions concerning
pictorial-symbolic intuition and its relationship to the sensory (pp.
194 ff). The result of those investigations led to the crucial charac-
teristic of geometric intuition: the possibility in principie of an
alteration of attitude in such a way that the sensible moments of
presentation of geometric state of affairs can be interpretad at any
time as depictions of morphological forms of the sensibly intuited
world of things.
While initially this determination of geometric intuitability was
related purely to individual geometric formations as such, now it
acquires its affirmation and support from the topology of normal
space. Since this is identical to the topology of the natural space of
objects, the "intuitability of geometry" signifies the pictorial, yet
pictorially-symbolic comprehensibility of geometric states of affairs
through the congruent connexus of euclidean normal space and
pictorial space. Each extensive manifold can function as a pictorial
space for the wherein of "depictions": the objective three-dimen-
sional space (the space of stereometric formations). or one that can
be regarded as pictorial surface ("sign-plane"). Topologically, both
are identical with the ideal normal space.
The failure to fulfill one or the other or both conditions, which are
not completely independent from each other, results in remarkable
deviations from the type of intuitability characterized.
Let us consider the case in which the condition of connectedness
between the ideal space and the pictorial space are fulfilled, but in
which the geometric state of affairs can no longer be presented
pictorially. The uniqueness of this case results from the limited
possibilities that are open for pictorial space with respect to its
dimensionality. Everything that is meant as pictorial-symbolic must
evidently be bound to two or three dimensions if an alteration of
attitude is to become possible with direct sensory perception of the
image as morphological "depiction." Such limitation lies in the
maximum three-dimensionality of the sensibly intuited world.
Formations higher than the three-dimensional cannot become
accessible in any kind of pictorial-symbolic intuition. The fre-
quently discussed question concerning the intuitability of multi-
254 Mathematical Space
dimensional geometry-and this also has to do with euclidean
geometry-must be emphatically answered in the negative if one
does not wish to burden the concept of intuitability with confusing
equivocations.
The manner in which n-dimensional formations (n> 3) are given
is so complex that even the most sophisticated analysis would not
venture to describe it fully and adequately. The mere suggestion of
an "extension" of geometry beyond the "boundaries of intuition"
(where by intuitability we understand only that within three-
dimensional geometry) leads no further into the problem. It does not
characterize what is specific in the nonintuitability of n-dimensional
euclidean formation in contrast, for example, to the nonintuitability
of noneuclidean geometries, which are completely distinct from the
topologically normal Sn-geometry. This differentiation, if it is given
at all to awareness, has not previously been an object of investiga-
tion. We shall attempt to explicate it only to the extent necessary for
the clarification of our problem.
It was tacitly assumed that normal space means the three-dimen-
sional euclidean space. The concept of normalcy obtained from the
metrics and topology of this space allows its extension toward Sn. It
too is of the topological type of the open plane, and its metrics is
determined by the euclidean movement group with the invariance of
the line segment.
But what does "line segment" and "movement" mean in more
than three dimensions? If for S
3
these meanings would result
through ideatively making them independent of any phenomena
belonging to the space of sensory intuition and through merely
carrying over their relations into the ideal domain, then their
conceptual apprehension in Sn would still require new moments of
apprehension. The latter would transcend the meanings of S
3
, and
while presupposing the meanings acquired in S
3
, would at the same
time transform them through the new meanings.
This would be the case if it were not for the fact that the
mathematical data of S
3
are already subtended by specific significa-
tions that are not first motivated by the constitution of S
4
They do
not appear in the plan as completely new moments; rather, they are
already co-constituted in what is given in S
3
and merely comprise
the exclusive sense-bestowing moments of 8
4
What is meant here
are the significations that enter into geometry universally with the
analytic-algebraic method. As already shown, this geometry com-
pletely disregards the number of dimensions and subsumes its work
purely under algebraic-formal criteria. Though it is capable of
Euclidean Normal Space 255
pictorial-symbolic illustration, in and of itself it does not require the
possibility of such illustrations.
Yet insofar as it is not merely a vectorial algebra but a vectorially
and algebraically pursued analytic geometry, the tendency of this
endeavor is to make it capable of a spatial interpretation such that its
vectorial equations must become conceivable as geometric states of
affairs. Thus it happens that even higher-dimensional geometry
speaks of "lines," "planes," "distances," and "movements"; it
speaks in concepts belonging originally to S
3
While the meaning-
content is the same in both, only in S
3
is it capable of symbolic-
intuitive fulfillment. The claim to fulfillment remains a predisposi-
tion even with the extension of these geometric concepts toward
higher-dimensioned spaces. A more thorough analysis of the sym-
bolic capacity of intuition would be able to show that even in these
spaces the originally pictorially perceivable moments are never
completely abandoned, but rather persist as conceptual roots. Yet
the latter cling only to concepts of objects taken in isolation ("line,"
"plane," etc.). In operating with them, in the apprehension of their
given positional relationships, this residuum is dissolved. Whether
aplane intersects another plane in Sn, or is touched by a specific line,
is not ascertained with the mathematically required stringency
offered exclusively by the calculus of determinants; every attempt to
establish the plausibility of algebraic results through subsequent
illustration also fails.
This specific nonintuitability of n-dimensional geometry is inher-
ent in the fact that it is a purely formal "extension" of the geometry
of S
3
Here the concept of extension is to be conceived purely
additively. That is why a converse reduction to three of the number
of vectorial components presenting the fundamental formations of
this geometry will result in three-dimensional euclidean geometry.
What is nonintuitable is thus present here not in sorne complex of
properties that are characteristic of the basic concepts of geometry
that resist illustration, but rather simply in the supercession of the
number of dimensions-if one may say so, in the numerical "exag-
geration" of the basic geometrical relationships, which in their own
right are never genuinely determinable numerically and thus cannot
be completely absorbed into algebraic relationships. Hence the
ground must be sought asto why, despite the complete topological
congruence between mathematical and possible pictorial space, the
possibility of a pictorial-intuitive fulfillment is completely absent
here.
It is quite otherwise with the case in which, according to the above
256 Mathematical Space
established division, the nonintuitability of a geometry is deter-
mined by the fact that the connexus of the ideal space deviates from
that of pictorial space, even if its number of dimensions corresponds
with or is less in number than that of the space of intuition.
In addition, a distinction must be made between spaces in which
the metric is identical with S
3
and those that are both topologically
and metrically abnormal.
The investigation of these spatial types must be added. Concerning
the question of their nonintuitability, we are limiting ourselves at the
present to the fundamentals.
That in particular this nonintuitability is constituted differently
for non-euclidean geometry than for n-dimensional euclidean geom-
etry is obvious from the mere fact that the latter is directly translated
into intuitable s3 geometry by its reduction to n = 3, while
noneuclidean geometry remains non-intuitable even within the
three-dimensional domain. Even for two dimensions the determina-
tion is difficult. Here non-euclidean geometry works as a theory of
surfaces, as a geometry of curved surfaces-of this later.
As a theory of curved surfaces it should be no less intuitable than
the euclidean theory of "surfaces," i.e., the normal geometry of
planes, since the curved is intuitable in the same sense as the
plane-in space. The meaning of this being-in contains a two-
pronged problem. A surface that is meant as "curved" obtains its
conceptual sense from the co-presence of the plane-structured space.
lt will be seen as a curved surface "embedded in" this space. Thus
a surface geometry with more than two dimensions cannot be
intuitable as long as it relates in any way to the surrounding
three-dimensional space, as is the case, for example, with the
elementary spherical trigonometry.
Note that the question here has no bearing on whether the
mathematical mastery of such curved spaces requires the inclusion
of a surrounding space. There is a widely spread false conception
that a curved mathematical space could "be" only "in" another
space of a higher dimension.
131
That it cannot be represented
otherwise is incontrovertible. Yet if its being consisted in nothing
other than its constitution through acts of representation of the
131. Thus N. Hartmann (3) has missed the problem of "ideal spaces," as
well as the problem of application of geometry, through such mistaken
notions. He obviously allowed himself to be led by the common and naive
notion; any attempt at phenomenological analysis of mathematical meaning-
giving is completely lacking in his "phenomenology."
Euclidean Normal Space 257
natural consciousness, then it would not be a specific mathematical
space. The manner and mode in which it can be represented does not
touch upon what emerges here with particular succinctness as the
genuinely mathematical: its complete freedom from representation
and independence from intuition. It owes its existence solely to
algorithmic thinking, whose basis of being is in no wise located in an
encompassing space of a higher dimension "surrounding" it, or in
which it is "embedded"-all these are only metaphors of a
consciousness requiring intuition and attached to the conception of.
space as a container.
Apparently it is a new discovery that curved surfaces can not only
be mastered from the standpoint of euclidean geometry, but that for
their own part they are spaces with characteristic geometric regularity
and "non-euclidean" relationships of measure deviating completely
from the normal. The usual geometry was first constructed as a gen-
uine theory of surfaces with Gauss, but only after the infinitesimal
calculus provided it with an analytic method. It was only with the
.conception of surfaces developed there that the relationships of mea-
sure attained their extension toward the higher-dimensional space of
non-euclidean geometry, in which those surfaces, as structural ele-
ments, play a role analogous to the plane in the euclidean Sn.
This will be discussed in greater detail in the next section. The
following chapter will briefly consider a mathematical type of space
metrically akin to euclidean space, yet revealing different relation-
ships of connectedness.
Cha pter Three
Euclidean Spaces with Topological
Anomalies
1. Extension of the Mathematical Concept of Space
On the basis of its constitutive moments of meaning, the mathe-
matical concept of space allows specific extensions. Thus the sense
of the mathematical concept of movement and its purely rational and
algebraically formulatable determinations reveals the fact that with
it-in contrast to the space of bodily movement-the ties to three-
dimensionality are abolished. They are not only abolished in favor of
higher-dimensional manifolds; an anologous "downward" exten-
sion also becomes permissible.
At first sight, the application of the concept of space to surfaces
may seem to be terminologically superfluous, if not nonsensical.
What sense can one attach toa space in space? Yet once again it must
be recalled that the decision concerning the being and sense of
mathematical spaces does not stem from the conception and mean-
ings of natural space-consciousness, but rather solely from the non-
contradictory construction of extensive manifolds in mathematical
consciousness. Accordingly, space is any manifold whose unique-
ness could be characterized by geometrical movements of its forma-
tions without regard to dimensional relationships. More precisely,
each mathematical movement group opens a mathematical space
independently of the dimensional number of what is moved. Even
surfaces, as two-dimensional spaces, fall under its concept. Whether
and to what extent they relate to a surrounding space of a higher
dimension is to be discussed later. The necessity, and the manner
and mode of such a relationship, can be meaningfully discussed only
when the use of the mathematical concept of space for surfaces no
longer appears as a merely analogous carrying over of its sense onto
surfaces-when it will instead be clearly seen that surface manifolds
are accessible toa mathematical treatment devoid of any mathemat-
258
Euclidean Spaces with Topological Anomalies 259
ical reference to a surrounding space. This would lead to a concep-
tion of such surfaces that would completely exclude their "surface"
character and grasp them as independent mathematical spaces.
The following investigation will consider the structure and metr-
ics of such spaces. Here our interest has nothing in common with the
special sciences, even if in what follows we cannot avoid having
greater recourse to mathematical states of affairs. Nevertheless, the
latter serve as clues for the noetic clarification of the subjective
activities that correlate to them; thus they are considered only to the
extent that they are conducive to an insight into intentional struc-
tures of acts of consciousness constitutive for their existence.
Following the preceding arrangements, the investigation begins
with a class of surfaces that on metric grounds still belong to the
euclidean manifolds; yet due to their topological anomaly, they are
akin to the non-euclidean spaces and lead across the threshold to
non-euclidean geometry.
2. Clifford-Klein Spaces
In 1873, Clifford discovered a peculiarly structured surface that in
many respects played a role in elliptical space comparable to the role
played by the plane in a normal (parabolic) space. Its mathematical
qualities were first investigated more precisely by F. Klein.
1
3
2
It is a
matter of a ruled surface of a second degree with specific properties.
Euclidean metrics is valid in it and its movement group includes a
genuine sub group of translations. While metrically it is of the type
of a plane, what is remarkable about this surface is that it is a closed
surface. Although its degree of curvature is zero, it contains a finite
surface area,133 This surface shows that vanishing curvature and
open infinity of space do not necessarily possess equivalent deter-
minations.
These peculiarities result from abnormal situs of the surface. As a
132. F. Klein (3). For more detail concerning the geometry of this form of
space, see also W. Killing (1), H. Hopf, and F. Ltibell (1)-(3).
133. More precisely, the metric congruence with the euclidean plane is
not valid with respect to the whole surface; topological anomalies are also
observable metrically. (This has to do with the multiple connectedness of
the surface.) Thus, for example, in contrast to the euclidean plane, where
each simply connected part-e.g., a line segment-can be moved in a
threefold infinite manner, here it can be moved only in a twofold manner
over the surface as a whole. Translations are normal, but rotation results in
deviations that are related to the various lengths of the geodesic.
260 Mathematical Space
particular ring surface, it is multiply connected. lts connexus can be
clarified in the following way. If one bends a rhombus into a ring
surface, so that corresponding locations of opposite sides coincide
with one another, then one obtains a formation from which one can
obtain a continuous one-to-one mapping of the Clifford surface. (The
mapping is not congruent; it is attainable only by an elastic mem-
brane.) Conversely, the Clifford surface can be transformed into a
simply connected surface: one starts with a point and severs the
surface lengthwise into two, so that they are simply bordered and
simply connected. This surface can be developed from a rhombus.
As can be shown mathematically, the Clifford surface investigated
by Klein presents a very limited special case of Clifford-Klein spaces.
W. Killing started with the peculiarities of this surface, showing that
it can be transported as a whole only by special movements. This
characteristic turned out to be appropriate for a series of surfaces or
spaces also having a non-vanishing (positively or negatively con-
stant) degree of curvature. Moreover, the aspect of generation of such
surfaces was permitted only by a specific mathematical lawfulness
of movement and of characterization; likewise for its construction
for more than two dimensions.
134
In contrast, very special types of
cases can be adduced precisely for the Clifford surfaces. Such would
be the surfaces that could always be developed-as in the case of the
Clifford surface-from the rhombus, from the parallelogram, or from
parallel strips. The formation developed from the latter becomes the
intuitable surface of a cylinder. lt is a one-sided surface and permits
a congruent development on the euclidean plane. The cylinder
should be regarded as a particular limit case of Clifford-Klein space.
If one considers an additional surface of a specific type ( double
surface), then it is obvious, as F. Klein has shown, that all forms of
space of this kind are exhausted. At the same time the topologically
abnormal euclidean spatial forms are exhausted.
3. Clifford-Klein Spaces as Euclidean Normal Space.
Founding Relationships
The specific topology of Clifford-Klein spaces requires that their
existence is not independent; rather, these spaces are dependent in
a determinate way on euclidean normal space. This dependence is
not to be understood geometrically; the mathematical treatment of
134. See W. Killing (1); the designation "Clifford-Klein spaces"- which
can be abreviated as C-K spaces-originates with him (pp. 257/58).
Euclidean Spaces with Topological Anomalies 261
such spaces does not require reference to a surrounding space with
a normal euclidean structure. Rather, it is ontological in kind and is
appropriate to the sense of existence of these spaces. They presup-
pose euclidean normal space, insofar as the mathematically consti-
tutive determinations which comprise their mode of being obtain
the fundamental meaning in euclidean normal space, from which
they then can first be meaningfully transposed into a Clifford-Klein
space. In accordance with their meaning, such concepts as a finite
surface area, "closed" surface, etc., imply the open-infinite space of
euclidean geometry, in and for which their precise mathematical
meanings were originally valid. What those spaces are, in what sense
being can be predicated . of them, depends on the mathematical
conception of space. This relates back to a specific geometric event
which, as a movement of certain fundamental geometric formations,
comprises space in the mathematical sense.
This relationship of dependence of Clifford-Klein space on
euclidean normal space is one-sided and not reversible. This means
that the euclidean normal space must "already" be befare the
Clifford-Klein spaces can claim existence in the specific geometric
sense. Their meanings are founded in normal space. This is obvious
on the assumption that euclidean normal space is the more primor-
dial mathematical space, having its immediate foundation in the
natural space of objects.
In view of this new form of space, it is possible to discuss the
question as to whether perhaps the space of the natural conscious-
ness of objects is already of such a structure that a Clifford-Klein
space could be regarded as a mathematical-ideative presentation
derived directly from the space of objects.
W. Killing maintains that it is basically possible for a Clifford-
Klein structure to be compatible with our spatial experience; there is
no compelling reason for us to suppose that with the movement of a
partial domain, the movement of the whole is necessitated, partic-
ularly since our empirical access is limited to bodily movements of
relatively short distances.
This question is discussed in detail by O. Becker.t35 Obviously
motivated by the presentation that E. Klein gives specifically to the
concerns of the Clifford surfaces, Becker begins with the mappability
135. See W. Killing (1), p. 258; see O. Becker (1), 12 B, 15. Becker limits
himself specifically to the first Clifford space mentioned here, which, dueto
its vanishing degree of curvature, must have motivated the question under
discussion.
262 Mathematical Space
of these surfaces onto open infinite space. If they can be developed
from a rhombus, then it is possible to think them as displaced to a
rhombus network that completely covers the euclidean plane. It
would represent all possible mathematical states of affairs of Clif-
ford's surfaces as periodically and infinitely repeatable. This would
be similar to the type of periodicity of the functional value of an
elliptical function. Correspondingly, the three dimensional space of
objects would be thought as articulated in accordance with rhom-
boids, such that all spatial events would have to repeat themselves.
periodically. At the same time, Becker reveals the "immense para-
dox" present in this structure as a spatial structure of material
things. It would not only be a space in which all physical events take
place, but would also have to contain the corporeally constituted
subject who apprehends space! But this would mean the abolition of
the irrepeatable uniqueness of one's own being in favor of a mere
exemplification of one's self, which would have to be thought
simultaneously with infinite frequency in each parcel of Clifford-
Klein space. This is logical suicide best relegated to the arbitrary
play of fantasy, although in an ontological reflection it can be
thought without becoming non-sensical.
According to Becker, this interpretation can be violated by elevat-
ing the distinctive topological structure of euclidean normal space to
transcendental understanding. If periodicity appears as a new rule in
the case of a Clifford-Klein structure, then there emerges a limitation
for spatial events, and this is incompatible with space as a
principium individuationis. The sale connexus that allows freedom
in space to events is of the type of open-endlessness of euclidean
normal space; it does not completely exclude periodicity, but does
not dictate it.
Despite the acceptability of the results concerning the distinctive
character of euclidean normal space, the way by which they were
attained is not without problems. The question whether the space of
material things, inclusive of the lived bodies of the space conceiving
subjects-i.e., the "actual" space-must topologically have the
normal euclidean structure, or whether it could also be the Clifford-
ian conception, cannot be decided in favor of the first, at least not
with arguments in which the distinctive structure. of euclidean space
is already presupposed. Yet this already tacitly functions as a
premise in the process of exposition of the consequences that the
Clifford-Klein structure should possess for spatial events; the latter
are not related at all to the Clifford-Klein surface itself, but rather to
the formations of development in the open-infinite (!) plane. For
Euclidean Spaces with Topological Anomalies 263
such development to be possible (for thought), the open-infinite
plane must be conceived ahead of time. And the sole motivation to
engage in such development is merely the need of a "euclidean
being" to visualize the Cliffordian relationships in a rough manner-
at least through means that topologically speaking are not without
dangers; they intrude into the situs of the surface and tend to
dissolve its ambiguities.
Finally, insofar as they cover the total open plane, the notion of
the infinite repetition of such development appears to be
problematic. Apart from the fact that the Cliffordian surface is not
identical with such a covering, the closedness of the surface falls
victim to a reinterpretation in intuition that has the greatest
relevance for the question posed above. lt is precisely the
closedness of the surface that suggests its infinite development and
can demonstrate in this manner the "recurrence" of events on the
surface. Yet there characteristically appears in the formations of
development a simultaneous repetition of infinitely many (in
relation to singular portions of congruently located) points. Thus
the repetition on the surface is represented as a temporal process of
"return" to the same point. Hence it should be noted . that the
presentation of surface events as well as their development involves
kinematic means, which in both cases imply the apprehension of
time as a linear and open-endless continuum. Assuming this, it
follows that the "periodicity" of events in a Clifford-Klein structure
is not just a result of a reinterpretation. After all, the phenomena of
returning to the same place within the surface at different temporal
points is subtended by their simultaneous recurrence at different
points of the formation of development.
If one wishes to demonstrate the priority of the connexus of the
normal euclidean plane, one cannot have recourse to periodicity as
a new law that does not appear in the Cliffordian surface itself. Even
Clifford-Klein space, as a principium individuationis, achieves the
same as does euclidean space and could not be simply rejected a
limine as a space of real things.
The conception that actual space could be of Clifford-Klein
structure may contain conditions that remain unacceptable to the
extent that 1 find myself in this space as a corporeal being. lt was
already shown that in the natural consciousness of space, the open
endlessness is not a simple factual datum of consciousness that in a
given case could be modified at will. Rather, its ultimate foundations
are to be discovered in the sense-founding intentionality of a
corporeal consciousness-the fulfillment of which can only be
264 Mathematical Space
guaranteed by the possibility in principie of an infinite progression.
Indeed, a Clifford-Klein structure of space would not completely
limit the possibilities of movement of a corporeal being, yet the
conditions of a perpetual-although infinitely repeated-return to
the same place would be contrary to the sense of the "progression"
in principie of movement "into" infinity. In principie the "aimless-
ness" of movement in a Clifford-Klein space, which is indeed further
movement but not a continuous forward movement in the precise
terminological sense, would not correspond to the sense-structure of
the intentionality of consciousness. In any case, even these last
clarifications presuppose open-endless time. All considerations con-
cerning the possibility or impossibility of topologically abnormal
structures for actual space are essentially and irrevocably tied to a
being that even in the conceptuality employed already reveals that it
does not find itself in the space in question, but can only attempt to
think itself into it-and that it can make such an attempt rests on the
basis of a conception of space (and time) that from the very outset is
determined by open-endless extension.
SECTION THREE
Non-Euclidean Spaces
Chapter One
Fundamental Questions of
Non-Euclidean Geornetry
1. The Parallel Postulate. Historical Origin and Development
As a mathematical discipline, non-euclidean geometry is conceiv-
able essentially from two sides. One was developed in the works of
Cayley and Klein from the viewpoint of the theory of parallels; it
remains clase to the elementary synthetic method. The other, the
analytic and infinitesimal-geometric, owes its establishment and
development to the achievements of Riemann. Although the latter is
mathematically more significant, the first cannot be left out of
consideration, on factual grounds. It permits the recognition of the
leading historical trends more clearly than Riemann's theory does,
and it has retained the oldest heritage of mathematical thought. In its
light, the controversia! parallel postulate first becomes understand-
able in its specific significance as well as in its important changes
which at the same time introduced a fundamental transformation of
the concept of the axiom.
The theory of parallels has its historical roots in the Elements of
Euclid. The proposition that is usually designated as the euclidean
parallel postulate--namely, that a given line possesses exactly one
parallel through a point lying outside of it-is held to be the most
outstanding fundamental presupposition of euclidean geometry.
136
136. We write "Euclidean" to designate the geometry that was constructed
265
266 Mathematical Space
Yet its unquestionability was already touched by Euclid's commen-
tator Proclus, who made it into a controversia! point of mathematical
discussion.
What is significant for our problem is that the parallel postulate
formulated in the above manner is not to be found in Euclid-,-there
is no "parallel" axiom or postulate (aitema) explicitly given by
Euclid. Apart from a nominal definition of parallelism, given in his
first book of his Elements as definition 23, there is under Euclid's
postulates at the fifth place the following proposition: if a straight
line intersects two others so that the sum of the inner angles of the
same side is less than two right angles, both straight lines intersect,
on the side of the angles.1
3
7
The existence and uniqueness of parallels is originally a very
specific consequence of this statement. It results from the basic
postulation of univocal relationships of ordering and, which was
most consequential for the subsequent discussions, from the presup-
position of the "existence" of the point of intersection which
certainly might have been obvious for Euclid.
Yet the incorporation of the proposition in question into Euclid's
system provokes sorne immediate reflections. Among the remainder
of his postulates (aitemata) the fifth-and in another respect the
fourth, concerned with the equality of all right angles-assumes a
special position. While the first three postulates require the possi-
bility of determnate constructive operations, and while this possi-
bility simultaneously establishes with certainty the geometrical
existence of specific formations, the postulate character (eitestho) of
the (fourth and) fifth postulates is truly distinct.
1
3B Indeed, two
in Euclid's Elements; in contrast, "euclidean" designates a geometric
domain that is to be delimited and characterized by the axiom named,
primarily with the emergence of the non-euclidean problematics. Thus
taken very strictly, Euclid knew no "euclidean" geometry.
137. Euclid, Elements I, Book 1, postulate 5.
138. In a striking way both are distinct from the rest even in their
grammatical form; both have something in common requiring that some-
thing "is." According to O. Becker (6), p. 214, the fourth postulate has the
character of a theorem and is to be seen as a consequence of th!l univocity of
the extension of a line segment. For the precision of the fifth postulate
Euclid requires that the original unity of the fourth and fifth postulates
should not be excluded. Becker holds that it is apparently possible to prove
that the fifth postula te was formulated befare Euclid without angles and that
only the convergence of straight lines constituted its content, while the
fourth postulate was still redundant.
Fundamental Questions of Non-Euclidean Geornetry 267
converging straight lines "require" a point of intersection-precisely
seen, the latter constitutes an existential staternent, yet the point of
intersection is rnerely stated but is not guaranteed by a process of
constructive requirernents! After all, the intersection of straight lines
discussed here is not an operation in the sense of the rernaining
postulates, but rather a consequence of an operation, narnely the
extension of straight lines, the possibility of which is guaranteed by
the second postulate. Thus the question of the "existence" of this
point of intersection differs from that of the remaining geometric
formations. It is not defined through construction but merely as-
serted; it is not generated constructively, which means at the same
time that it cannot be proven.
This remained a problem for later times. The controversy concern-
ing Euclid's fifth postulate was unleashed for the first time by
Proclus. For its recognition, he demands mathematical proof. With-
out it he holds that the Euclidean claim for the existence of the point
of intersection is a mere probability.139
As subsequent axiomatic investigations taught us, the demand for
proof by Proclus placed an inappropriate demand on geometric
science. Yet what is eminently positive in his critique of Euclid,
which should have led to an important discovery is the detaching of
the characteristic point of intersection of two straight lines from their
convergence. Proclus does not challenge the convergence, but the
point of intersection-it is possible that two straight lines moving
toward one another do not intersect, even if they come into an
unlimited proximity to each other. Proclus supports this claim by
the fact that with other kinds of lines there is convergence, but at the
same time it is only an asymptotic approach. He stresses this
possibility also for convergent straight lines. Their point of intersec-
tion is not universally contested: what is brought out is that two
somewhat converging straight lines could run symptotically.
Factually, this includes the possible existence of many "parallels"
for the straight line through one and the same point, and strictly in
accordance with the euclidean definition of parallelism. After all,
the latter is not given positively with the aid of a concept of
equidistance, but rather employs the state of affairs of the non-
intersection of straight lines (in the same plane).
1
4
138. For the discussion of the concept of aiterna in ancient rnathernatics,
see also K. v. Fritz (1).
139. Proclus, pp. 191, 16-193, 7.
140. What is rneant by Euclid is obviously the (subsequently designated)
268 Mathematical Space
In his commentary on Euclid, Proclus does not offer any clue that
would suggest that he was fully aware of the extent of the thoughts
that had led him to the threshold of a non-euclidean (hyperbolic)
geometry. What later became influential was primarily only his
demand for the proof of the controversia! postulate.
It is not the concern of our investigation to trace the attempts at
proof in detail; a brief indication of the structure of such proofs
should suffice. In all cases they started from the sum of two right
angles in the triangle, a state of affairs that is equivalent in meaning
to the questionable postulate. Usually they employed an indirect
process of proof: they started with a premise opposite to the o'ne to
be proven and attempted to obtain a contradiction. At best, the result
led to the insight that the parallel postulate is replaceable by
equivalent statements, as long as what is to be proven does not
already function as a more or less hidden premise in these
"proofs."
141
At this point we should mention only the remarkable
attempts of G. Sacceri (1735), who suggested two further possibilities
besides the euclidean postulate. In tracing their consequences, he
followed the notion that he could lead both to absurdity and in this
manner indirectly prove the euclidean postulate. (Sacceri speaks of
the "hypothesis of the obtuse angle" and "hypothesis of the acute
angle." By this he means a hypothetical postulation of right triangles
in which both angles placed in opposition to the base angle-with
both being equal to one another on symmetrical grounds-taken
together are either greater or smaller than two right angles. The
refutation of the first hypothesis succeeded under the assumption of
the open-endless extension of straight lines.) That Sacceri succeeded
in refuting the one hypothesis-and this under a completely speci-
fied presupposition-and that in contrast his attempts in the other
failed, must have been the impulse that led J. H. Lambert to a new
reworking of the problem. With him the discussion of the euclidean
parallel postulate assumed its decisive turn.
Lambert no longer attempts either to prove orto disprove Euclid's
postulate, but shows how parallellines follow from his own theory.
affine plane. For antiquity, the ideal elements of a projective plane were
completely outside of the mode of observation.
141. Specifically, the proposition of the sum of angles in the triangle is
equivalent in meaning to the Euclidean axiom of parallelism. This may have
been one motive among others that led again and again to attempts to prove
this postulate, especially since the opposite of that proposition is demon-
strable.
Fundamental Questions of Non-Euclidean Geometry 269
He does not manipulate and reconstruct the problem as posed but
modifies the question. What happens in geometry if one admits both
of Sacceri's hypotheses? Lambert was the first to gain an insight into
the noncontradictoriness of three "geometries," among which one
preserves the euclidean parallel postulate and in subsequent termi-
nology is "euclidean," while the other two are non-euclidean.
Lambert deserves the credit for founding a new method in geometry
and preparing the ground for the subsequent theory of proof. Since
then, the provability or non-provability of a mathematical funda-
mental presupposition is no longer the deciding factor concerning its
axiomatic character-rather, the assumption of an axiom, or the
change of an axiomatic system, determines the form of geometry.
This leads to the abandonment of the traditional conception of the
essence of what is geometric. And this is incorporated by Lambert in
his theory of parallels. The full insight into the logical nature of the
problem should be seen as his most significant contribution. It is
astounding how modern is his demand that in geometry "one must
abstract from the conception of things" and must proceed "purely
symbolically" (this means in our terminology signitive-symbol-
ically) in geometric proofs.
The controversy over the demonstrability of the parallel axiom
ends with the insight into its superfluity. While Euclid's concern
was to lead back from the multitude of individual geometric states of
affairs to a limited number of fundamental principies, the new
method proceeded in the reverse direction: beginning with a limited
number of fundamental presuppositions, it constructed new geomet-
ric manifolds and built them in accordance with strict procedures.
The way was thereby not only freed for work in non-euclidean
geometry by Gauss, Bolyai, and Lobachevski, Helmholtz and
Riemann; it also prepared the ground for subsequent foundational
investigations carried on in the second half of the 19th century by
Pasch and systematically developed to its full extent by Hilbert.
2. Constitutive Problems of the Parallel Postulate
The inner-mathematical interconnection of a given axiomatic
system and geometry is only of limited interest for this investigation.
Of decisive importance for us is solely the mode and manner in
which the new geometric manifolds are constituted in mathematical
thinking. Thus at the outset there is a question concerning the way
in which the euclidean parallel postulate is to be understood and the
genuine locus of the constructive conditions of its content. Obvi-
270 Mathematical Space
ously, it is not a question that touches upon the axiornatic or
postulate character as such. This would be the case with the
mathernatical view. Rather, it airns simply at a phenornenological
tracing of its content back to those presuppositions frorn which the
latter would be understood as the intentional correlate of specific
mathernatical attributions of rneaning.
The postulate in the sense of Euclid obviously has the following
rneaning: the point of intersection in question exists in each case of
convergence of two straight lines, even when it no longer falls
within the pictorial-syrnbolic presentation in the (lirnited) pictorial
plane. In this "no longer" the structure of the thought process
becornes visible. There is a tacit assurnption of the co-existence of
two states of affairs-the convergence of the straight lines and the
existence of an intersection-forrnation-simply because it is
evident in pictorial-syrnbolic intuition for all those cases in which
the point of intersection falls within the limited pictorial space,
given sufficiently "strong" convergence of the straight lines. In
these cases the point of intersection is trivially guaranteed through
construction. Correspondingly, a "weaker" convergence of the
straight lines can be compensated for in principie through the
extension of the pictorial space. From this it can be concluded
analogously that a point of intersection of two lines moving toward
one another must be generally present, when in "weaker"
convergence the point is removed farther and can no longer be
factually constructed but only assumed. Such an assurnption must
have been without question for Euclid. It was unthinkable for him
that the geometrical event could in principie occur otherwise in the
large rather than in the small, i.e, in a limited plane accessible to the
mathematical operations of those times. As one can see, Euclid's
parallel postulate-more clearly, his assumption of the existence of
the point of intersection-contains in its turn nothing else than a
sensible extension of a geometric event in a limited region of space
toward the totality of space. It is analogous to the extension that was
previously discussed for the limited space of intuition and the
natural space of objects.
To be sure, for Euclid, straight lines and point of intersection are
no longer things of a natural space of objects-nor are they on the
other hand conceived by him as formations of a "mathematical
space." To pose such a question of space would be inappropriate for
the geometry of antiquity. But what is significant for the euclidean
conceptual process is first the fact that here one begins with states of
affairs that lead to the genuine parallel postulate (the existence and
Fundamental Questions of Non-Euclidean Geometry 271
singularity of parallels through any point that lies outside any given
straight line) by a kind of consideration of limits, which unquestion-
ably moves toward the possibility that a limited domain of space is
capable of unlimited extension. This limit case is given when both
inner angles taken together result precisely in two right angles. And
secondly, what is remarkable is that this conceptual process begins
with geometric states of affairs that are not only accessible in
pictorial intuition but have the sale condition for the possibility of
geometric existence, for the geometry of that time, in such pictorial
symbolism. The existence was therefore bound to the medium of its
presentation, whose sign was the "plane". With respect to its
structure, the latter corresponds topologically to the natural space of
objects. In accordance with its origin, the euclidean claim under
discussion is bound to a geometry that can be carried out only in a
space structured by planes; aild the "euclidean space" subsequently
determined from it as a free geometric manifold can be such only if
it manifsts plane structure and maintains the same connexus as the
natural space of objects. Seen phenomenologically, this is the
structural uniqueness of euclidean space. Seen mathematically, it is
the fifth postulate of Euclid that expresses this uniqueness of
euclidean normal space. In arder not to dismiss this euclidean
presupposition of planarity as mathematically naive, as is so readily
done today, one must become cognizant of its truly deeper founda-
tions in arder to understand that it has long continued in the
mathematical tradition. It seems that not only line segments and
angles, but also lines and planes have their pre-scientific
signification-fulfillment in the space of intuition. Their specific
mathematical sense is first understandable only in terms of such
signification. These topological formations are, as scientific concep-
tions, grounded in particular in preceding achievements of the
subject as a corporeal being and his co-given functional constitution.
After all, the ray-like character of the visual function allows us to
understand the topological structure of the space of intuition in
terms of lines and planes (pp. 248 ff.). Geometry had neither reason
nor possibility to assume any other form apart from plane geometry
as long as its sense of existence originated from the mode and
manner of a constructive generation. This was more so since the
latter, in a most original sense, was still a production of its forma-
tions and remained an instrumental production. They related to
space only to the extent that the latter was already there as lived
space, which provided the objects "on" which those formations
reached conceptualization through a highly complex process of
272 Mathematical Space
achievement. Lived space was simply the condition for their spati-
ality and pictoriality.
With Sacceri the state of affairs is not yet different. Led at the
outset by the intent to guarantee Euclid's foundation for geometry
against all attacks on the fifth postulate, he too, like Euclid, begins
with the states of affairs in limited pictorial space and thinks that he
can then carry it over unto space in general. Since he tacitly assumes
the open-infinity of the line in his refutation of the hypothesis of the
obtuse angle, his attempted proof rests on the intuitive background
of the plane space.
With Lambert there appears a noteworthy turn.1
42
With the use of
the Saccerian hypothesis of the obtuse angle, he obtains a spherical
triangle, and thus relates his hypothesis to a spherical surface.
Certainly, the recognition that on the sphere there are triangles
whose sum of angles is greater than two right angles was not new.
Menelaus of Alexandria, a contemporary of Plutarch, had already
developed a geometry of the spherical triangle in analogy to Euclid's
theory of triangles, and his "spheric" belonged to the classical fund
of spherical geometry. It is all the more remarkable that many
centuries of controversy had passed concerning Euclid's parallel
postulate befare Lambert ended it with the obviously simple and
illuminating suggestion of spherical triangles.
The originality of Lambert's account would only be missed if one
wanted to see in it more than a renewed indication of long since
known geometric results or to take his incorporation of spheric-
trigonometric components into the discussion of parallelism as
nothing other than a mere justification of a geometric hypothesis for
a limited partial domain of geometric res.earch. What is decisively
new in Lambert's thought is rather a fundamental transformation of
a methodologically basic conception of geometry. He pointed to the
spherical surface not because it is a reservoir for non-euclidean
relationships, but solely because non-euclidean events can be
demonstrated from it. Lambert stresses the hypothesis of the obtuse
triangle not for the sake of the long since known spherical
trigonometry; rather, he wants to offer the same justification for his
hypothesis as is done for the euclidean parallel postulate. At the
same time, he appeals to the sphere as merely one domain of its
intuitive realization. Yet how little this has to do with the validity of
the hypothesis follows unambiguously from Lambert's remark that
142. An extensive appreciation of Lambert is given by P. Stii.ckel and F.
Engels.
Fundamental Questions of Non-Euclidean Geometry 273
the hypothesis of the acute angle would then be valid for an
"imaginary" spherical surface (what is meant here is a "spherical"
trigonometry for triangles with imaginary sides and real angles, ora
geometry on a "sphere" with an imaginary radius-a formation that
is no longer representable in any pictorial-symbolic sense).
All pictorial symbolism in geometry now functions merely as a
heuristic means. This is valid henceforth not only for the new
geometries, but also for the euclidean. The leading role is assumed
solely by the algebraic symbol, the pictureless sign for the spatial,
whose pictoriality has now become geometrically meaningless. This
strict abstention from grasping the spatial in a pictorial symbolism is
demanded by Larn.bert for the first time in all rigor for geometry.
Expressed positively, the consequent possibility to signify exhaus-
tively the spatial only in signs simultaneously abolishes for sign-
geometry the unbridgeable opposition established by real-spatial
and pictorially intuitable meaning: the opposition between plane
and non-plane surfaces, between open-endless and endless-closed
space. Speaking purely phenomenally, this abolition is visible in the
simple fact that signitive symbolization discards any ties to a plane
medium of presentation. For a geometry that is purely signitive and
constructive in the specifically modern sense, it is in principie
irrelevant whether one operates in a medium of open-endless
extension or with a form closed upon itself. The plane of signs, in its
constitution as plane, is as irrelevant for the characterization and
meaning of a sign as is the color of the chalk used for the process of
proof; and the latter is not changed in the least in its symbols and
steps of procedure if it is somehow transferred onto a curved
( one-sided) surface.
This is apparently trivial-and yet it simultaneously shows a pro-
foundly penetrating differentiation of performance-between picto-
rial symbolism and sign symbolism in geometry from a new si de. The
first demands that the plane of signs on which something is de-
signated, and the plane indicated in the signs, should be topologically
congruous with the formations-euclidean elementary geometry is
not synthetically constructable on a spherical surface. In contrast,
sign symbolism permits the sign-plane, on which it is merely signi-
fied, to deviate topologically from the normal plane. In principie, it
is thinkable that for an analytic geometry no other real-spatial for-
mations would be needed for its production apart from surfaces of a
higher order. Nothing could either take away its existence projected
merely signitively in a conforma! geometry or somehow modify the
mathematical sense of its existential propositions.
274 Mathematical Space
This does not mean, however, that with the signitive construction
of geometrical manifolds, "plane" and "surface" become indiscrim-
inately equivalent, although they become formations of equal rank.
To consider geometrically a (curved) surface as a plane means to
modify its conceptual sense. This not only concerns the surface
purely as such "in space," but also results in the constitution of
completely new types of geometric spaces. For these, the surface
then plays a fundamental role corresponding to that of the plane as
a structural element in euclidean spaces. In those spaces, however,
a specific geometry is consequently valid which already dominates
the surfaces. This geometry is non-euclidean insofar as the underly-
ing surfaces do not fulfill Euclid's parallel postulate.
But of course the geometry discovered by Lambert is not yet purely
non-euclidean in the precise sense. The spherical surfaces on which
the new insights are demonstrated are with him not yet strictly
:rion-euclidean surfaces, but rather are still conceived as surfaces
with spherical trigonometry. Even Lambert's geometry is factually
still a trigonometry of surfaces in space, i.e., still in the euclidean-
structured three-dimensional manifold. But it is not yet a geometry
of a completely new space. At that time, the manifold did not even
enter consideration as more than two-dimensional. In the conceptual
sense of this geometfy at that time it could not be arbitrarily
extended dimensionally. Lambert's conception of surfaces basically
hides the fundamental novelty of a structure of thought that is
nevertheless quite clearly present in its inceptual form.
A surface can first be called "space," however, when it turns out
to be a manifold with its own rules for geometric events, and also
when its mathematical conception is able to exhibit a reference back
to mathematical embedment in space. This offers, then, the possi-
bility to vary the number of dimensions. As Riemann's work
subsequently demonstrated, such formulation was, in accordance
with Lambert's demand, purely symbolic. Lambert's formulation
first lays clown this requirement, although it could only fulfill it
partially itself.
We are now confronted with a question: which founding intercon-
nections should be traced in arder to reveal the ontological
problematics of the non-euclidean manifolds and in arder to grasp
them as new unities of sense of the space-positing consciousness-
the same consciousness that does not simply surrender the
euclidean presuppositions, but rather first grasps them in their
genuine and fundamental meaning with the new manifolds.
Chapter Two
Foundational Problems of Hyperbolic
Geometry
1. On the Metrics of Hyperbolic Geometry
This tapie is particular! y important for that form of non-euclidean
geometry which, as the geometry of the acute angle, was related by
Lambert to the imaginary spherical surfaces. Yet this relationship
can no longer be simply represented on a surface like real spherical
geometry. Rather, it is presented as nothing other than a purely
algebraic relationship, which incorporates imaginary numerical
quantities as constructed analogues to the usual spherical
trigonometry.
Seen historically, it is the first geometry that was constructed
precisely as non-euclidean; it is the so-called hyperbolic geometry.
Proclus had already pointed out the possibility of numerous paral-
lels going through one point. For this geometry Gauss discovered the
functional dependence of the sum of the angles of closed figures on
their surface content; in this way he obtained the limit case of the
"infinitely large triangle" with the vanishing sum of angles. In
contrast, the euclidean relationships are dominated by sufficiently
small triangles.
The framework of this investigation does not require detailed
exposition of the mathematical states of affairs of the new geometry.
At present they are significant only insofar as they reveal the
existence of phenomena that lead to grasping the question of the
founding of hyperbolic geometry in sensibly intuited factors, and
thus to the question of their intuitability. While at first glance the
latter may be denied, a closer look reveals sorne interesting problems
whose specific attraction consists in their remarkable interrelation-
ships with the space of intuition.
To trace them we can use the work of Felix Klein as a point of
departure. By following the suggested direction and by accepting the
275
276 Mathematical Space
Cayleyan projective determination of measure, he was able to
incorporate the hyperbolic as well as the elliptical geometries as
special cases of general projective geometry.1
43
This incorporation
became possible because this geometry is related to a specific
fundamental surface of the second degree. In accordance with this
foundation, the differences among geometries are formed solely in
terms of the type of such fundamental surfaces. In other words, the
specificity of a given geometry results solely from the specificity of
such surface.
Thus to pursue geometry, it is necessary to conceive of the line
segment as the basic formation of all metrics. While Euclid
designates it nominally as the "distance between two points." it is
in fact determined as an invariant property of euclidean motion. In
a very specific sense this latter characteristic is lost in non-
euclidean geometry. Through the new determination of measure,
i.e., the projective, the invariance of a line segment is replaced by
the invariance of the dual relationship of four line segments; in the
surface the distance between two points results in a mutual
relationship of location of four points (two of which can possibly be
imaginary). Klein's main contribution to the determination of
measure in hyperbolic geometry consisted of his definition of the
logarithm of such a dual relationship as "distance" between two
points in the fundamental surface.
At first glance this appears to be logical nonsense. To define
"distance" or "line segment" by using four line segments for the
definition obviously contradicts all rules of proper definition. Yet in
a positive sense this definition contains the specific characteristic of
a modification of the meaning given here to the concept of line
segment. In the original euclidean sense this is a presupposition for
the possibility of the structuration of dual relationships. It contains
a categorical constitution of a unity of a higher arder; it is a
specifically graded synthesis of the kind in which general "mathe-
matical expressions" are constituted and which in this case implies
the mathematically exact pre-constitution of the line-segment as
such. As a condition for any determination of measure, the line
segment is presupposed by all measures. Yet in no wise does it
regulate the activity of measuring. Its simple "marking off" is, as a
linear operation a+ a+ a+ a ... merely one among many possibilities
of its application. Under the operative aspect, it presents itself only
143. F. Klein (1), (2), concerning the "Erlangen Program" see especially
the formulation in (2) in Math. Ann. 43.
Foundationa1 Problems of Hyperbolic Geometry 277
presents itself only as the primordial form of any measuring, which
nonetheless can be varied in principie. Thus the dual relationship,
in accordance with its geoni.etric sense, is nothing other than such a
variation in the manipulation of line segments for the purpose of
new measuring. It does not define the concept of the line segment-
which is already presupposed here-but rather it prescribes a
particular manner and mode of proceeding with line segments.
While this procedure results in a new determination of measure,
in the present case the projective, it seems to lead to the notion that
the concept of line segments "acquired" a meaning that is different
from the euclidean. Yet despite all this, the concept is "defined"
with the aid of line segments, leading toan apparent contradiction in
definition. The contradiction nevertheless dissolves immediately if
we consider that we are not concerned with the usual definition of
a concept and the establishment of an exact objective meaning, but
rather with an establishment of a procedure that prescribes what
must occur with the euclidean line segments in non-euclidean
geometry. The new concept of the line segment, in distinction to the
original, is nota concept of an object, but an operative concept. In its
new conceptual sense it is constituted with, and on the basis of, an
originary giving of meaning toa line segment in the acts of operating
and proceeding with it in accord with the mathematical prescript of
its dual relationship, i.e., of its logarithm.
This kind of prescript may astound a non mathematician. Ulti-
mately what is normative for its sense is only the totality of the
algorithm; it can become discerned only in terms of its algorithm.
What motivates this prescript can be shown suggestively only with
respect to what it can accomplish. The new determination of
measure turns out to be so designed that it allows the use not only of
distance in hyperbolic and elliptical geometry but also of the
euclidean distance (of "parabolic" geometry) inasmuch as even the
latter can be formulated as a logarithm of a dual relationship. It thus
constitutes for the new metric conception a characteristically math-
ematical mode of generalization. The non-euclidean distance is
present in the projective determination of measure as a result of
mathematical generalization of the euclidean determination of mea-
sure; the latter in turn is nothing else in its projective aspect than a
particular specification of the non-euclidean. The nature of this
specification is more closely determined by the mathematical struc-
ture of the logarithmic function. This means, in addition, that this
non-euclidean geometry is euclidean in its smallest parts.
Furthermore, it is interesting that the suggested mathematical
278 Mathematical Space
states of affairs have a specific type of intuitability for hyperbolic
geometry. Indeed, it is fundamentally distinct from that of euclidean
geometry, yet even if in a very modified sense, it shares with it a
certain picturability of its geometric states of affairs. After all,
hyperbolic geometry, as a special case of universal projective geom-
etry, has its uniqueness in that its charecteristic surfaces consist
purely of real points. Following the far reaching implications that
the fundamental surface of hyperbolic geometry is real-valued, and
that it can be pictorially-symbolically presentiated, Klein created for
the geometry related to it the well-known "model." In the sequel it
will be sketched in its basic outlines for two-dimensional hyperbolic
geometry.
144
2. The Kleinian Model. Phenomenological Analysis of the
Model Conception
For hyperbolic geometry the fundamental surface becomes a
fundamental conic section that is intersected by any line of the
(euclidean!) plane at two real or two imaginary points (the latter
means: pictorial-symbolically "not"). It can be conceived as a circle
or an ellipse. Klein's notion was to limit the manifold of points
within the conic section and to ascribe to them a logarithmic
determination of measure. The logarithmic function has a character-
istic such that in this determination of measure, the conic section
turns out to be the locus of "infinitely distant" points; the conic
section itself lies at a "logarithmically infinite" distance. The conic
section chords function as "lines" of this geometry. The interior
angles have real numerical measure. The multitude of parallellines
through a point, has in this way a simple intuitive fulfillment:
"parallel" are those straight lines that intersect at "infinity," i.e., at
a point of the conic section (or outside of it). The conic section itself,
"atan infinite distance," is never reachable-because of its charac-
teristic interior determination of measure: a "line segment" of this
geometry, approached by a non-euclidean being that can only move
in the interior of this conic section, is constantly moved back in
logarithmic foreshortening or, measured in euclidean time, a being
144. We limit ourselves here to a model of F. Klein for hyperbolic
geometry, since we are not concerned with the enumeration of all model
conceptions for non-euclidean geometry, but only with the exposition of
what characterizes a model conception in geometry as such. Kelin's model
can be taken here as an example.
Foundational Problems of Hyperbolic Geometry 279
inside the fundamental conic section moves ever "more slowly"
toward the edge without ever reaching it. The interior of the conic
section functions as its plane of movement; this being cannot know
what is beyond the boundary of the conic section. (In an ideal
mathematical sense it would at the most be approachable, as would
be the infinitely distant line in euclidean geometry for a euclidean
being.) Since all interior distances are measured by logarithmic
relationships, there are only logarithmically determined triangle
sides. The sum of angles in the triangle varies with the size of the
surface content; for infinitely large triangles (in the euclidean for
inscribed triangles), the surface content is zero. The attempted
descriptive sketch of non-euclidean relationships indicated a signif-
icant equivocation. On one side, it touched upon an anthropomor-
phism readily used in the sciences. A "non-euclidean" being was
confronted by a "euclidean" ("external") being; what for the latter is
the interior of a conic section, for the former is aplane; what for one
is a conic section chord, for the other is a line; what for one is a conic
section, for the other is infinitely distant.
It is necessary to illuminate the type of understanding obtaining
between these two beings. To do so we must divest the specific
structure with which we are concerned, the structure of the "as," of
its anthropomorphic cloak in arder to make it conceivable in terms
of the giving of meaning of the one subject who alone is capable of
grasping this description-that subject. which is not only the
"euclidean" and at the same time the "non-euclidean" being, but
also the subject concerned with such exposition of this mutual
understanding.
The subject himself is a euclidean being in the sense that he has
already opened euclidean space and has mastered it mathemati-
cally-and has done so on the basis of the natural spatial
apprehension founded ultimately in his modes of corporeal
comportment and on the basis of the mathematically constitutive
achievements of ideation, symbolization, and formalization in the
sense presented above. Remaining at this level of constitution,
living at the same time in the euclidean space of action, this subject
experiences the conic section, chords, etc., primarily in their
original geometric meaning-bestowing, i.e., as formations of the
euclidean plane. Through the calculus of the new determination of
measure, motivated by the mastery of non-euclidean relationships,
there occurs then the change in meaning of those originally
conceived euclidean formations, in their pictorially symbolized
"self," into a new sense-bestowing of them "as ... " The change is
280 Mathematical Space
such that a fixed and all pervasive relationship of coordination is
established for all singular formations between the original and the
modified content of meaning. This relationship of coordination is
determined by the new regulations of measurement of line
segments. This is always taken into account when a chord is taken
as a line, a conic section "as" a totality of infinitely distant points,
its interior "as" a new total plane, etc., and the resultant sense of the
original concepts is in its turn understood as meaningful. The
subject already has the calculus in readiness, in the pictorial-
symbolic projection of those formations, when he approaches the
pictorial plane prepared to illustrate, with its help, what has been
achieved through purely algebraic analysis. Hence the fiction of a
hyperbolic being becomes redundant for its interpretation of
non-euclidean relationships; its relationship to euclidean being is
understood as a relationship between two different bestowings of
meaning to the pictorial-symbolic formations in one and the same
consciousness "befare" and "after" the calculus. In a certain way
this regulates a translation of meanings without allowing the
translation to have a corresponding pictorial image-the figures,
taken purely as such, rernain "euclidean," i.e., they are first posited
by virtue of their euclidean rneaning. They are preserved latently
when consciousness "lives" in re-signification of these forrnations
through a new calculus of rneasure.
What is pictured and what is meant enter into a unique
relationship of tension in hyperbolic geornetry. The signified here is
not the original geometric rneaning of the pictorial syrnbol, but
rather is constructed frorn it; the latter rernains in the background of
consciousness in the acts of non-euclidean signifying, which use an
algebraic calculus. The latter, however, is not itself pictorial, but
syrnbolic in a purely signitive sense. This rneans that the pictorial
syrnbolism of the euclidean plane is inappropriate and pictorially
inadequate for what is genuinely meant in it. This, however, is
no longer accessible via pictorial presentation in any other rnan-
ner than the one sketched above, and thus it cannot be intuited
in the previously maintained sense of this term. In place of the
simple pictorial-symbolic intuition in euclidean geornetry, there
enters a "model conception" as a mediating kind of syrnbolization
which-in distinction to pictorial symbolisrn in the usual geornetric
sense-requires an insertion of an algorithrn. This is achieved in
such a way that the model must be independently constituted ahead
of time in arder to fulfill its rneaning in sorne way in a sensible
medium.
Foundational Problems of Hyperbolic Geometry 281
3. Hyperbolic Geometry and the Space of Intuition
If the model conception of hyperbolic geometry is characterized as
an algorithmically mediated type of symbolic intuition, it stands nev-
ertheless in a particularly close association to the natural space of
intuition. lt gains this affinity from the fact that it suggests a geometry
of a limited plane ( of the interior of the conic section) in which the
euclidean event is seen, soto speak, in a distorting mirror. Here there
is obviously a remarkable analogy to the relationship between the
natural space of objects and the singular space of intuition. This too
is limited horizonally and, with. regard to the relationships of par-
allelism, it specifically points to a noticeable similarity to hyperbolic
geometry. Even within the "domain" of sensory intuition there are
many "parallels" intersecting one and the same phenomenal point
(compare p. 122), and even a being walking into the distance walks
virtually ever "more slowly" toward the horizon without ever reach-
ing it. In a specific domain of nearness, euclidean relationships seem
to rule, while in analogy to hyperbolic geometry, the deviations are
greater toward the border. The formal analogy is so noticeable that it
threatens to reverse completely our previous results: it is not the
euclidean, but rather the non-euclidean-hyperbolic geometry that
apparently must be regarded as the more primordial, insofar as its
founding strata are already immediately contained in the horizonally
limited space of intuition. Thus the question emerges: should a more
careful phenomenological analysis of the mathematization of space
have led not primarily to the euclidean but rather to the hyperbolic
geometry? In our previous presentation were we perhaps led simply
by the factual development of history and not by systematic problems
of constitution?
If this were the case, if the historical point of view had clearly been
the leading one, and if euclidean geometry had thus "systemati-
cally" assumed the first place in our investigation only due to its
historical priority, then a critical revision of the preceding analyses
would be in order. If hyperbolic geometry is more fundamental and
is rooted more deeply in the corporeal-sensible event and its spatial
world than is the euclidean, then it could appropriately claim full
intuitability in the sense of the previously meant pictorial symbol-
ism; in contrast, euclidean geometry should only claim intuitability
of the model type.
But this obviously conflicts with the composition of the phenom-
ena in the mathematizing consciousness. The difficulty in the
illustration of hyperbolic geometry was demonstrably located in the
282 Mathematical Space
requirement of a special calculus that first had to mediate the
relationship between the pictorial symbolic intuition of geometric
formations and the meaning-giving of something given in the figure
"as" newly meant. Thus here the illustration could only function as
a model. But this means that this merely conditional type of
intuitability, which is mediated in a very involved way, rests upon
the ontic priority of normal euclidean relationships. For it was
euclidean geometry that provided the basis for the new determina-
Han of measure and at the same time grounded the phenomenal
preponderance of normal pictorial symbolism in the hyperbolic
model.
It remains true that hyperbolic geometry is conceivable as the
geometry of a limited mathematical space. It also has a type of
morphological correspondence to a limited space of intuition. Yet
this correspondence does not have a character of a founding rela-
tionship. Hyperbolic geometry cannot be interpreted as a mathemat-
ical idealization of the intuitable morphological world of things in
the perspectiva! space of intuition, analogously to the idealization of
the elementary euclidean formations (see pp. 184 ff.).
In addition, it ought to be recalled what truly determines the space
of intuition as such and what role accrues to its objectivity for the
intentional constitution of geometrical phenomena. Although in its
being the singular space of intuition is relative to a corporeal here,
nonetheless it turned out to be at the same time not merely a spatial
"part" of an infinite space, but rather a manner and mode in which
the latter can existas such and, as a whole, for a corporeal subject.
The one homogeneous space is constantly co-present in the singular
perspectiva! space of intuition and, in this mode of co-presence, it is
already constitutive for the phenomena of perspectivity and of
horizon itself. It is only on this basis that we comprehend such
cognitive characters as size constancy, changelessness of the form of
a thing in motion, and unlimited possibility of motion. Hence the
mathematical sense-bestowing to be gained from these for
euclidean geometry is already understandable from the space of
intuition. Strictly speaking, it is not first the homogeneous space of
objects, but rather its co-presence in the space of intuition, that
allows the latter to comprise the basis for euclidean space. Similarly,
the analysis of the foundations of the euclidean phenomena must
have recourse to the sensibly intuited data of perspectiva! space.
The correspondence between intuitive data and hyperbolic geom-
etry must be understood more precisely. Strictly speaking, it is not
the space of intuition in which a type of hyperbolic regularity is to
Foundational Problems of Hyperbolic Geometry 283
be found, but rather the viSual space. Only visual space-not the
space of intuition-allows one to speak strictly of a boundary
without beyond, and a multitude of "parallels" through a single
point is a visual and not intuited (they are intuited precisely as
parallels; (see p. 122). Visual space is determined as an abstract
moment of the space of intuition. It is acquired from the latter by
excluding the co-positing of "the" space and by reducing the full
thing of intuition to the merely visible. The visual objectivity does
not present any originary phenomenal data, but constitutes itself as
such in its "thetic" modes by disregarding certain co-given factors
in the primordially intuited data.
The relationship between the space of intuition and visual space
has unique consequences for geometry. Just as visual space is
phenomenologically inaccessible without the space of intuition,
likewise the hyperbolic space cannot be conceived without the
euclidean. Just as visual space is grasped first "from the side of" the
space of intuition, so also the constitution of hyperbolic objectivity
in mathematics first succeeds by going through euclidean normal
space. Nevertheless, this formulation contains an easily misleading
analogy. As was shown, euclidean normal space is founded in the
space of intuition, indeed founded directly (p. 252). Since the former
founds the space of hyperbolic geometry, hyperbolic geometry is
also founded mediately in the space of intuition.
However, visual space does not "found" the space of intuition;
rather, it is its dependent constitutive moment and comes into
"view" as independent only abstractively, through a process of
exclusion. That it is possible to find in it certain correspondences to
hyperbolic geometry indicates that the pure lawfulness of vision,
which participates in the full sensory intuition, might possibly
underlie non-euclidean relationships and that hyperbolic geometry
is applicable to it. But this does not mean that the latter is founded
in visual space. "Application" is nota founding with a reverse sign.
The founded and the founding have a necessary intentional and
essential interconnection. The former builds itself in acts "upon" the
latter, and requires the latter in arder to be. An "application" has an
entirely different structure. Something to be applied is in principie
free in contrast to that to which it is being applied. Indeed, the latter
is required to allow the former "to find" application, yet the
possibility of such finding application does not at all touch that
which exists for itself without and beyond the application. A
mathematical objectivity preserves its pure mathematical sense of
existence even if it is never and nowhere applicable. Hyperbolic
284 Mathematical Space
geornetry is in its own right, even if the laws of vision do not obey
it. Its sense of existence is related solely to the required rnathernat-
ical rneaning-giving, and it is only frorn there that it can be
introduced rneaningfully into the discussion concerning its applica-
bility in visual space. That the latter rnight possibly "be" non-
euclidean is a clairn whose ontological sense is not irnrnediately
obvious. It rnust rather be explicated in its own sense and delirnited
precisely against the ideal sense of being of geornetrical rnanifolds.
This leads in general to questions of applicability whose problernat-
ics rnust rernain excluded frorn this investigation.
Chapter Three
Riemann 's Geometry
1. Riemann's Point of Departure. The Metric Fundamental Form
Riemann's work places the problem of non-euclidean geometry on
a new foundation. The novelty of his point of departure, based on
infinitesimal calculus, is not only internal to mathematics. His point
of inception is also relevant to our problem. He advances a com-
pletely determinate conception of space as such, subtended by its
mathematical mastery. Riemann's program-designed to understand
the world in the large from the infinitely small-offers in contrast to
the traditional treatment of geometry (in the sense of the theory of
parallels) a new methodological situation. The infinitesimal geomet-
ric treatment of space is decidedly based on this methodologically
fundamental requirement: to limit oneself in all geometrical propo-
sitions strictly to infinitesimal domains and to refrain from any
propositions concerning the structure of space in the large.1
45
While
traditional euclidean geometry, starting with geometrical states of
affairs in a limited domain of space, simply transfers them to space
without seeing any special problems in such transfer, with Riemann
there is an explicit requirement to abstain from such judgements.
This is elevated to a fundamental methodological requirement.
What is unique in this is not only that one can successfully
investigate the previously recognized mathematical spaces as a
whole in light of new mathematics, but above all that Riemann's
method creates t ~ e possibility for constructively projecting new
kinds of mathematical spatial structures.
The mathematical relationships obtaining between euclidean and
Riemannian geometry may at first sight blur their decisive differ-
145. The question concerning space as a whole is called by Riemann an
"idle question"; for him only the infinitesimally small deserves scientific
meaning. See B. Riemann, III. 3.
285
286 Mathematical Space
ence. If it is the case that what is generally admitted as characteristic
of Riemannian geometry is also found in euclidean geometry-
though "only" in the infinitesimally small-then obviously we must
come to an appropriate conception and explication precisely of this
"infinitesimally small." Since questions of the mathematical contin-
uum are introduced here, subsequently we shall take a position with
regard to it. Here we must primarily note that the notion of
"Pythagoras in the small" does not at all touch the core of the
Riemannian point of departure.
Indeed, the euclidean determinations of measure in the small hold
true of Riemannian geometry; what this means in the mathematical
formal language is that the Pythagorean theorem assumes the alge-
braic form of a differential expression. Y et this is a form of the
Pythagorean proposition that first results, for the Riemannian deter-
mination of measure, from a further and more general principie that
is not to be found at the beginning of Pythagorean deliberations. The
infinitesimal turn of the Pythagorean proposition concerning the
line segment is primarily a specification of a more general mathe-
matical expression, the so-called metric fundamental form.
Euclid's geometry, and the conception of space as a manifold of
planes co-posited for it, is insufficient for its attainment. The general
foundation for its investigation is seen by Riemann in Gauss's theory
of curved surfaces.
It is methodologically decisive that in contrast to the elementary
geometric conceptions, these surfaces are no longer used for the
additional demonstration of specific geometric states of affairs, still
retaining the residual conception that they themselves are forma-
tions "in" euclidean space; rather, such surfaces are understood at
the outset as properly geometrical manifolds.
As independent spaces, they are determined essentially by a
pertinent coordination. They are no longer related to the three
coordinate directions of S
3
Rather, each determination of location is
oriented exclusively in accordance with two proper coordinates or
parameter curves chosen in the surface. In its mathematical presen-
tation, the surface as a whole assumes a vectoral form, where the
surface vector is a variable of two parameters. Correspondingly, the
length of a line segment is no longer determined within the surface
as a spatial curve with three functions but solely from the two
surface parameters. The mathematical expression resulting in this
manner for the length of a line segment in the surface is the metric
fundamental form. Taken generally, this contains a result of a
specific differential expression for the element of length. This means
Riemann's Geometry 287
that all values in the surface are dependent on certain tensor
magnitudes which, as partial derivations of the surface vector, are
constituted in accordance with its parameters. They are constant
functions of location and determine the geometrical event of the
surfaces, determine the "metric fiel d." The geometry on a surface is
determined univocally through these tensor magnitudes; the surface
is already characterized by them. In turn, each mathematical expres-
sion of the form just sketched can be conceived as a metric
fundamental form, and can be taken as a constitutive element of a
new geometric space.
At present it is not our concern that the fundamental form
suggested is not yet the most general: nonetheless, the Riemannian
point of departure already contains two remarkable characteristics.
One is the infinitesimal turn of geometric problems, and the other is
the positive definiteness of a quadratic form for the element of
length. The first provides the new methodological means for the
complete treatment of non-euclidean problems as such; through the
second characteristic, the method is contentually bound to a number
of geometric spaces which, as "Riemannian spaces" in a narrower
sense, differentiate themselves from the still more general non-
euclidean spaces through this characteristic of the positive definite-
ness of their fundamental form.
2. Riemannian Spaces. Brief Mathematical Characterization
A geometry constituted by the determination of length of the type
of a positive-definite quadratic differential form is characterized first
of all by its ability to subsume euclidean geometry as a special case--
i.e., a Riemannian geometry shifts to euclidean geometry when the
tensor magnitude deploys its fundamental form in constant num-
bers. Even with the preservation of its complete tensor character-
more clearly, in its general determination as position functions to
which accrue distinct values in accordance with position and
surface orientation-it is not without relationship to euclidean
geometry. Even in this general case the euclidean metric is valid for
the "neighborhood of any point." Here the accent must be placed
completely on the "neighboring" characteristic, since the infinites-
imal point of departure strictly forbids the transference of mathe-
matical relationships found in the small to a larger domain.
Certainly, it is peculiar that Riemann himself has not followed out
this point of departure with those consequences that he had
programatically demanded. Riemann begins with a determination of
288 Mathematical Space
measure. that includes the position-independence of length and
allows, or at least does not exclude, the possibility of comparing
distances between lengths. Strictly speaking, this is counter to his
fundamental principie. According to Riemann, this euclidean resid-
uum results in a determination of measure that is characterized by
the property of integrability of lengths. Accordingly, a mathematical
space is determined as Riemannian space if each line segment in it
can be moved as an invariant along any chosen path; the numerical
measure of a line segment is also here independent of any path.1
46
The congruent transfer of line segments is again to be strictly taken
as infinitesimal. This demand contains what is perhaps one of the
146. Riemann's point of departure is capable of extension in two
respects. When the metrics of a surface accrues solely to the metric
fundamental form, and 'furthermore, when each expression of this kind can
be interpreted as a metric fundamental form to be chosen as a constructional
element of new mathematical spaces, there is nothing that prevents the
surrender of the limitation by the positive definiteness of the fundamental
form still maintained by Riemann and thus the construction of spaces that in
sorne characteristics would deviate from Riemannian spaces. More signifi-
cant than this deviation is another, a generalization of Riemannian thoughts
assumed by H. Weyl. It presents a consistent and more fully developed
extension of Riemann's concepts insofar as the infinitesimal mode of
consideration is applied primarily not to vectors but to lengths. Thus a
possibility is given where not only directions, but also lengths are not
integrable and hence are path-independent. In these "general metric spaces"
the characteristics of a vector of a given length at one point cannot decide
about its characteristics at another place. To enable a congruent transfer, the
neighbourhood of the point must be so adjudicated that the length of the
vector in this neighbourhood can be maintained. The metric structure of this
space also contains the specific gauging that enables exact propositions
about the transfer of lengths. Mathematically, this structure is presented as
a linear form for the establishment of the so-called curvature of the second
kind; this is a measure for the changes of length experienced by a line
segment in the course of displacement on a closed path. Like the first
fundamental form, this one is also invariant in contrast to coordinate
transformations. Thus, in conjunction with the former, it constitutes the
characterization of the relevant metric space. The Riemannian spaces
distinguish themselves from the Weylean generalization through the iden-
tical vanishing of the curvature of the second kind, the "normal gauging."
According to Weyl, it is necessary and sufficient for a space to be a
Riemannian space. A more detailed discussion of Weylean spaces is
superfluous within the framework of our problem. Concerning their math-
ematical treatment, see H. Weyl (2).
Riemann's Geometry 289
main difficulties in the process of following out Riemannian
thought. Primarily, it Gan merely suggest to us what the differential-
geometric point of departure brings about and what deviations from
euclidean geometry it entails. Although the euclidean residue in
Riemannian geometry preserves the relationships of line segments in
the euclidean sense, there nevertheless appear notable differences as
soon as one envisages movements not for line segments but for
vectors. Apart from length, they also possess a direction, resulting in
a number of specifications.
Indeed, as a consequence of its metrics, Riemannian space eo ipso
possesses an affine connection. This means that any given vector is
displaceable parallel to itself to any given place of the Riemannian
space, although naturally this displacement is an infinitesimal
displacement "from point to point." The displacement is ruled by
specific transformations that correspond in priQ.ciple to the displace-
ment transformations in euclidean space. On the basis of the
infinitesimal metric point of departure, however, "parallel" dis-
placement of a vector results in a change of its direction "from point
to point," and this change of direction is dependent on the path of
displacement. If one thinks of a vector (ora "localized vector group")
in Riemannian space as having a closed path, returning to its point
of departure through iteration of its displacement, it retains an
invariant length for the length of this path. Yet-pictorial-
symbolically-it appears in general as not having returned to the
point of departure, and this deviation must be different in accor-
dance with the path of displacement.
This non-integrability of directions in Riemannian space is, de-
spite its original opposition to the conception of parallelism, any-
thing but a paradox. Logically it follows as a completely univocal
result of the methodologically fundamental proposition in virtue of
which the surfaces are characterized by the previously mentioned
tensor magnitudes, which enter into the differential expressions as
fundamental magnitudes regulating the infinitesimal partial dis-
placements.
In contrast, for a surface in which the tensor magnitudes are
constant, these differential expressions provide a vector transfer
independently of paths. Here after traversing the closed path, the
localized vector group returns to its point of departure. This is
precisely the case with aplane surface, as well as with all surfaces
that can be developed from a plane. A space structured in accor-
dance with planes turns out, from here on, to be necessarily
euclidean in the sense that the components of affine connection
290 Mathematical Space
vanish in them. Through the integral iteration of infinitesimal
displacement there results the finite transformations of the total
space, leading back to the fundamental meaning of the concept of
parallelism in the sense of elementary geometry.
Among the possible paths of displacement of a vector, in Rieman-
nian space, only sorne have a characteristic that results in parallel
displacement in the euclidean sense. Thus with the closedness of the
path of displacement, the parallel displacement leads the vector
back to its position of departure.
This is also a geometrical relationship that regulates the general
transformation of displacement without transcending the framework
of the generallaws of displacement of such surfaces. The apparent
agreement with the euclidean meaning of parallelism lies in the
geometric specificity of such paths. In other words, it is inherent in
the properties of the differential laws of displacement in such
surfaces, that for certain paths, they require parallelism in the
euclidean sense with mathematical necessity, and such a path
attains a distinction above all possible paths precisely through this
parallelism.14
7
Such distinctive paths are given with the affine
connection of the surface and the behavior of the localized vector
group. They represent this connection; they are characteristic for the
given surface as a whole.t4a
Their totality constitutes for each surface a doubly infinite mani-
fold. This is similar to a manifold built from lines of a plane. The
latter analogy is supported by the fact that those curved paths
distinguished as "geodesic lines" possess in geometric measure an
important property, enabling the comparison with the straight lines
of plane-structured space. Geodesic lines are curves of stationary
lengths; this means that for surface metrics they represent the
shortest connection between two surface points. The shortest con-
147. It is also to be characterized from the side of S
3
and respectively
from the surrounding space of the surfaces, since under this aspect it is at
the same time a spatial curve. Its specificity then consists in the incidence
of the surface normal and the principal normal of the curves with respect to
all of their points.
148. More precisely, it can be said that what solely conditions the kind
of occurrence of such distinctive paths is not the metric, but the affine
connection. They are concerned only with the differentiation of directions
and not with vectors. This state of affairs becomes significant, for example,
for the concept of the geodesic for surfaces with non-integrable length
transference (see Riemann-Weylean spaces, note 146).
Riemann's Geometry 291
nection for two points of the euclidean plane is offered by a
connecting straight line. Analogously, one can proceed from one
point of a surface to another in the shortest way only by taking the
geodesic path. In a certain sense, the geodesic lines represent the
straight lines in a non-euclidean space. The analogical use of the
concept of a straight line is allowed even in a further respect. In the
euclidean plane, each path connecting two points, apart from the
shortest, is a curved path14
9
and the shortest is distinguished
everywhere by vanishing curvature. It is possible to define a concept
of curvature for non-euclidean spaces in such a way that with
respect to them, the geodesic lines have the characteristic corre-
sponding to the straight lines. Thus the "geodesic curvature" of a
surface curve vanishes only for those curves that are geodesic, while
any other connecting line in the surface, in relationship to this
concept of curvature and in contrast to the geodesic, is "curved." At
the same time, the concept of geodesic curvature clearly delimits the
conception of "spatial curve" from that of the other, the "surface
curve," for the sorne formation: whereas the geodesic-except for the
straight lines of the plane space-appears in general as curved,
insofar as it is conceived as a line of a "surface in space," it is a
geodesic line of the surface itself. Considered in ~ non-euclidean
sense, it is of such construction that its geodesic curvature is zero.
But, for example, the circumferences of a "sphere in space" are
circles of a determnate radius of curvature and never straight lines.
Yet as lines of the spherical surface and as exclusively related to it,
they are geodesic lines with zero geodesic curvature, and in this
sense "straight lines" of the surface.
This results in a concept of curvature that contains completely
different mathematical meanings. The indicated differentiation of
meaning giving becomes possible only on the basis of two funda-
mentally distinct mathematical meanings of the surface itself. It will
become necessary to consider them once more when the frequently
discussed and often misunderstood theme of "space-curvature" is
taken up.
3. Curvature and "Curved Spaces"
The brief consideration of the fundamental metric form already
suggested the distinction in meaning between surfaces in space and
149. Connected series of lines {Polygonzgen} will be disregarded, prima-
rily beca use as connecting paths they are not too interesting mathematically,
since they are not differentiatable in all points.
292 Matbematical Space
surfaces as space. In the first conceptual sense, the surface remains
tied to the intuitive representation, and indeed it is essentially
representable only in this sense. It remains related to a spatial
embedding medium of a higher dimension and furthermore, if it is
taken as a curved space, its medium is a euclidean structure. As
already mentioned, any mathematical treatment of a surface in this
embedding space is to be conceived as euclidean.
It is otherwise when the mathematical relationship toa surround-
ing space is abolished. This is the case in the full development of
surface geometry, which must be seen as the genuine non-euclidean
geometry and which, in its turn toward the infinitesimal, constitutes
the essence of Riemannian geometry.
The decisive impetus for this conception of the surface as an
independent mathematical space lies in that the surface is provided
with its own coordination which, as two-dimensional, inheres in the
surface itself and allows possibilities of dimensional extension for
the mathematical treatment of surfaces-just as euclidean geometry
has the plane for a plane-structured space of any chosen dimen-
sions. The surface thereby becomes conceivable in principie as a
structural element of non-euclidean spaces of a higher dimension.
The geometric independence of such spaces is apparently counter
to the fact that they are concerned with curved manifolds. In accor-
dance with their sense, the latter obviously presuppose a surrounding
space of higher dimension and plane structure; it is only in contrast
to the latterthat they can be seen as "curved." Obviously, the concept
of curvature implies the meaning of planarity. Curvature is conceiv-
able essentially only as a "deviation from" the plane, while in con-
trast the opposite is not valid. The plane already revealed its funda-
mental meaning in its structural distinction in the space of intuition
which has its correspondence, on the subject's side, to the pregiven
structure of the laws of vision (pp. 249 ff.).
Geometry justifies this irreversible relationship insofar as it ap-
proaches the mathematical conception of curvature with methods
that determine the degree of curvature of a curve or a surface in terms
of the plane-structured space; the degree of curvature is seen as a
measure of the deviation from the plane relationship. Thus the
mathematical determinations of curvature by means of the radius of
curvature, the first variation of the are length, etc., presuppose in
principie the plane-structured surrounding space. Concepts such as
a constant or a variable curvature of a surface in this first and
simplest mode of mathematical apprehension have pictorial-
symbolic fulfillments of meaning. In rough morphological terms,
Riemann's Geometry 293
they designate the "regular" and the "irregular" way of being curved
relative to the plane-structured natural space of objects, for which
the degree of curvature is zero.
In all this, such unique apprehension of curvature would be
counter to the conception of a surface of two-dimensional proper-
manifold. lt can only be maintained as such if the mathematical
concept of curvature can successfully be determined in such a way
that the curvature would be conceivable purely in terms of the
surface relationships themselves. The decisive contribution of Gauss
is to have shown this possibility. His famous theorema egregium
ends with the proof that the position function determining the
curvature of a surface is determinable so1e1y from the tensor magni-
tude of the metric fundamental form through which the internal
measuring of surfaces is achieved completely and univocally with-
out recourse to an embedding medium. Thus the Gaussian curvature
is mathematically of equal significance with the internal relation-
ships of measure of a surface. Whether it is a constant ora position
function, changing from point to point at both surface parameters, is
determinable purely from the geometric event in the surface. In
anthropomorphic terms, a two-dimensional being existing only in
the surface would be in a position to determine the curvature of
space precisely and indeed solely from the relationships of measure
of its surface without necessarily having recourse to a surrounding
space.1
50
That this Gausian curvature is solely a so-called "inner" concern of
the surface can be demonstrated by a particular property of
150. Thus the Gaussian curvature would be capable of yielding a
principie of classification for the various classes of non-euclidean spaces.
Hence the spaces of constant Gaussian curvature are characterized by their
having (infinitesimal) congruent displacements, and the size of a geometric
formation in them exists independently of position. Spaces of this state of
affairs are distinguished as metrically homogeneous spaces. They include
the ("elliptical") spaces treated by Riemann as well as the hyperbolic and eo
ipso the (parabolic) euclidean space. From the totality of all possible
mathematical spaces, these "spherical surfaces" are the only ones that have
the distinction of being spaces of congruence. (The geometry of the sphere
must be taken into this division, though it is not identical with Riemann's
spherical geometry; rather, it is distinguishable from it, in the topological
sense, through its one-sidedness. The spherical space is two-sided in
three-dimensional space. The transition results from the topological identi- --'
fication of two diametrically opposite points. Like elliptical space, spherical
space is thus also homogeneous.
294 Mathematical Space
invariance. If one bends a surface (without stretching or distortion),
then the Gausian curvature does not change. Bending, as meant here,
is not a mathematical concept; rather, it is morphological. lt is
capable of sensible-intuitive fulfillment and can be found in any
possible occurrence of bending. In contrast, when in the terminology
of the discipline the Gaussian curvature is designated as bending
invariant, this does not give any mathematical precision to the
concept of bending; rather, this designation shows the sense of a
non-identifiability of curvature and bending. Morphologically, the
concept of bending is not at all differentiated from the concept
(equally morphologically related) of curvature. Whether a real object
cylindrically formed has a "curved" ora "bent" externa! surface is,
within the morphological domain of signification, a useless question
and merely a conflict about words. Only the mathematical domain
offers an exact differentiation between these concepts. And indeed,
each intuitive-morphological bending can in principie be grasped
with mathematical exactness-but not every such bending is a
curvature in the Gaussian sense. Thus in the example given, the
bending of the cylinder may be ascertained in natural intuition; the
Gaussian curvature nevertheless remains zero. Correspondingly, not
every mathematical meaning-giving of curvature can be represented
pictorial-symbolically as bending.
Even this most simple example warns us against our wanting to
encumber a mathematical concept with ah intuitive (pictorial-
symbolic) fulfillment of meaning. While its original conceptual
construction is founded in the morphological domain, its subse-
quent acquisition of meaning in the continuous pursuit of new
geometric problems has become removed from its intuitive founda-
tion. Intuition can no longer be fully sufficient to it, and any pictorial
symbolism must become inadequate. lt is only by preserving this
state of affairs that one escapes the erroneous opinion that curvature
could be had only for a spatial manifold that is "in" another
(plane-structured) space and is dependent on it, as if the concept of
curvature is countersensical for higher than two-dimensional forma-
tions. This concept signifies precisely that there are given specific
"inner" relationships of measure that are just as little bound to the
limits of dimensioning in the non-euclidean as the euclidean deter-
minations of measure are bound to the euclidean plane. As always in
the construction of concepts of geometry, here too we are concerned
with the sense-bestowal that is built primarily upon the original
foundation of meaning belonging to the lived spatiality capable of
pictorial representation. In mathematical consciousness, however,
Riemann's Geometry
295
the constitution of these sense-moments is motivated by a series of
viewpoints that are not at all specifically geometric viewpoints.
They not only lead to more precise and differentiated conceptual
dimensions, but also to conceptual transformations, which for their
part are most distinct in kind. In their totality they have their
ultimate legitimacy in specific requirements of universality of the
mathematical sciences. They are so conceived that they not only
transcend the particular, but also simultaneously contain the latter
as a special case.
It is with the aid of this conception of curvature that the discussion
of "space-curvature" is first accessible to meaningful exposition. If
any residuum of euclidean understanding is abolished in it, if the
understanding is presented in a strict non-euclidean sense so that it
mathematically masters the surface exclusively from within itself-
which means that at the same time it envisages its curvature as
nothing else than a specific inner structure of the measure of the
surface-then in principie there is the possibility of any dimensional
extension whatever of a two-dimensional non-euclidean manifold.
Seen formally, the ascent into multi-dimensional non-euclidean
spaces occurs in the same way as the euclidean; in both cases it
fundamentally means only an adjunction of further parameters
whose mastery is exclusively an analytic-algebraic concern.
In the infinitesimal basic structure of this calculus there is, at any
rate, the fact that non-euclidean spaces are not first non-intuitive for
higher dimensions. Rather, precisely to the extent that a two-
dimensional spae is taken to be non-euclidean, it already resists
intuition. The appropriation of concepts from euclidean elementary
geometry-length, vector, parallel displacement, etc.-may here
pretend to have a certain intuitability, just as what is expressed in
them is in fact originally capable of pictorial-symbolic illustration.
Yet this picturability vanishes as soon as these qualities are consid-
ered in terms of what happens with them in the non-euclidean. That
a line segment can change its length and that a parallel displaced
vector can change its direction, and that by returning to its point of
departure it does not in general return to its position, is really not
representable; this completely eradicates the intuitive fulfillment of
meaning of "length" and "parallel" and evokes a peculiar experi-
ence of conflict. An infinitesimal congruent transfer results ulti-
mately in incongruence, and an infinitesimal parallel displace-
ment-as soon as and precisely when, it is sufficiently reiterated-
finally leads to considerable differences of direction in the sense of
the original meaning of the concept of direction.
296 Mathematical Space
This has nothing to do with meanings that are "carried over" or
"transposed," or with a mere use of terms. Rather, the result seems
to have conceptual meanings, which in their core maintain an
apparent affinity to the euclidean. Yet this turns out to be an illusion
when understanding turns to geometric events in the non-euclidean.
The geometric event in these spaces, ruled purely by analytic
algebra, not only no "longer" leads to intuitive fulfillment-as
would be the case in higher-dimensioned euclidean geometry-but
rather simply opposes it. With its basic point of departure it disrupts
the lawfulness of the sense, and the construction of objects capable
of such fulfillment.
The discrepancy between what occurs here geometrically and
what direct intuition would anticipate occurring is grounded in the
circumstance that pictorial representation always essentially in-
eludes what Riemannian geometry excludes in its basic point of
departure. The former encompasses a specific finite domain whose
arder of magnitude far surpasses what is solely allowed by the
infinitesimal-geometric point of departure, and encompasses it si-
multaneously as a pre-given domain. In contrast, the differential-
geometric calculus opens out this domain successively, and indeed
through the infinitesimal iteration of geometric events "from point
to point"; it establishes it first of all in operations.
In addition, this discrepancy may be decisively grounded upon
the fact that what direct intuition takes to be a continuously
extended domain, differeiial geometry grasps in such a way that it
concurrently prescribes an algorithm for conceiving this continuum
itself. The usual mode of discourse claiming that something must
occur "from point to point," or from "infinitely contiguous points,"
hides a deeper problem which ultimately is aporetic in nature.
4. The Question of the Existence of the Mathematical Point
With the problem suggested, which at the same time touches u pon
the problem of the continuum, we are approaching the conclusion of
our investigation. Jt, reaches its limit, in a twofold sense, with the
question concerning the meaning of existence of the mathematical
point. First of all, this question no longer belongs in this domain.
Rather, its discussion compels philosophy to engage in pure analy-
sis. The problem of the continuum-as a problem of classification of
all irrational numbers, well-ordering, countability by means of
transfinite numbers, etc.-belongs ultimately to pure mathematics
and lies outside of the geometric questions. What is relevant for us
Riemann's Geometry
297
turns out to be merely the question of the intuitable continuum and
its rational conception and mastery by way of the algorithm.
Secondly, our investigation encounters methodological limits-
even in the suggested narrowing of the problem. They lie partially in
the uniqueness of this problem itself. Indeed, our investigation is
concerned with more than accepting a given objectivity offered by
mathematics as something existent and, in accordance with our
procedure, tracing it back to its constitutive origin of meaning. The
question still remains as to whether we are really dealing with
mathematically existing "objects." Certainly this is still controver-
sia! in the discipline of mathematics.
151
Yet only as such an object
would the point and the continuum be accessible to phenomenol-
ogy. While this attempt will be made in the concluding part, it can
be justified, if need be, by exhibiting the problem of the continuum
within the limits in which it is present within the framework of
geometry, and thus revealing the limits of the method used to date.
We revert back to the continuum in lived space. There it presents
no "problem"; it has a direct intuitive presence as a phenomenal
continuum (pp. 88 ff.). For lived space, the concept of the continu-
um had the meaning of denseness and non-interruption of spatial
parts, sorne of which are individuated as "places" of things. Yet
there is the constant concern with blurred, morphologically vague
articulations without sharp boundaries. Insofar as the region in the
space of action, and the phenomenal point of the space of intuition,
are topologically the ultimate elements of the respective spatiality,
they essentially escape precise localization. They are conceived as
the finite limits of the nesting of domains prior to the exact
determination of boundaries.
While for direct intuition they are everywhere dense, for the
mathematically thinking consciousness the process of nesting is
capable of progression that far surpasses what is sensibly conceiv-
able. Mathematics structures space further "inwardly" all the way to
the formation conceived as the mathematical point. This does not
avoid, but simply hides the problem.
lts attainment obviously requires a further iteration of the topo-
logical contraction and of "thinking" the nesting of domains "to
infinity." This process of nesting immediately runs up against a
blind alley. If the progressive iteration of nesting leads to increas-
ingly smaller spatiallocations, which can be conceived as contigu-
151. Concerning this problem in general, see the subtle analyses of C.
Becker (2), especially pp. 583-620.
298 Mathematical Space
ous and touching with locally adjacent boundaries, then the iteration
of this conception leads to a dilemma, especially if such nesting is
regarded as progressing to the point which, according to Euclid, "has
no parts." This difficulty leads directly to an aporia: how is a sum of
such points, regardless of how densely they may be conceived, to
constitute a continuous space? After all, a continuum is character-
ized by the fact that it must be divisible into continuous divisibles
and thus it cannot exist as a multitude of points, since points are
indivisible. Should a continuum be thought to consist of points, the
points themselves would have to be continuous; this is contradicted
by the concept of a point. This conception faces an inescapable
alternative: either it abandons the punctiformal conception of the
continuum in arder to do justice to the point, or it abandons the
continuum in favor of the point.
Yet the existence of the spatial continuum is not to be denied.
Unaffected by all the difficulties of its mathematical conception,
continuous extension is a phenomenal datum, indeed a "direct"
intuitive datum. It is founded primarily in the fact of continuous
motion. Only the attempts of its rational conception could be
condemned to failure. But this does not place the continuum itself
into question. What is in question is the existence of what is called
a "mathematical point." Not only is it in no sense an intuitive
datum, but even in mathematics its existence is burdened with all
sorts of problems. We shall discuss it here only as a question
concerning the extent to which its claim to existence can be given
the meaning that was previously attributed by us to geometric
existence in the sense of constructability.
The existential question of the mathematical point is thus prima-
rily a question of its access. Whether and in what sense "there is"
something like a mathematical point is to be decided solely from the
mode and manner of its conception. With respect to its treatment in
Euclid's Elements, the point apparently assumes a remarkable dual
position. If in the Euclidean sense the existence of a geometrical
formation is guaranteed through constructive activity having firmly
prescribed means, then the geometrical point cannot exist for it, at
least not in this sense. In Euclid the point is not guaranteed through
construction (something like a formation of intersection), but rather
is assumed as given in arder to guarantee the geometric formations
through construction.1s2
In contrast, there is the variously criticized pedantry-factually
152. Euclid, Elements, Book 1, postulates 1 and 2.
Riemann's Geometry 299
unjustified-with which Euclid constructs the irrational quantities
in Book X of his Elements, quantities that are conceived to be the
relationships of a line segment to a given initialline segment. The
explicit and detailed work that Euclid dedicates to every newly
introduced irrational relationship impressively shows that he does
not simply accept the existence of all irrational (quadratic) points or
numbers. Rather, he sees the necessity of securing their differently
comprised mode of being-megethe and not arithmoi-by a unique
way of construction. This construction of irrational numbers must
have been important for Euclid, since the obviously specific mode of
existence of these quantities could come to light only in it. For
Euclid, they "exist" only to the extent that they can be established
constructively, yet they do exist in this mode even if they are given
as unlimited (apeiron) from the standpoint of rational number
relationships.
It must not be overlooked that the irrationalities existing in
Euclidean geometry remain limited to the quadratic. Since for
Euclid the only existent formations are those that are attainable
constructively through circles and lines, then geometrically speak-
ing, only such points are acceptable as can be composed from
quadratic roots and the irrationalities consisting of them, and no
others. And yet there are these points in the sense that geometrical
existence had for antiquity. They are acquired from constructive
generation in precisely the prescribed way.
15
3 The limitation of the
constructive means of antiquity contains the fact that antiquity's
153. A point of division tripartitioning a circle's are thus do es not exist for
antiquity, even if intuitively it may be evident that there must be one. It
comes into existence later when a new means of construction was intro-
duced by means of conic sections (circles and lines in space). There is
nothing in Euclid concerning their application. The only exception to the
quadratic irrationalities in the geometry of antiquity is the number pi, which
is applied for cubings and rectifications. (All the problems of this kind at
that time were solely dependent upon pi). Yet pi here assumes a unique
place to the extent that this transcendent number is "introduced" as a
relationship between circle surface and square of radius, i.e., it is not
constructed with the pre-given means of construction, since pi cannot be
constructed with the means of antiquity. If a convergent approximate
construction were to be allowed as validation for the proof of existence, it
would have contradicted antiquity's basic conception of the essence of
mathematics since the existential proofs of antiquity are always concerned
with finite constructions. Strictly speaking, the number pi does not exist in
the sense of antiquity's mathematics.
300 Mathematical Space
definitions of irrationalities are not limited solely to quadratic roots,
but also that such definitions are finite in principie. The existence
of irrational quantities is ascertained through finite construction
and their entire sense of existence is exhausted by geometric
construction with the given implements of construction (compass
and straightedge) in a finite number of steps. Today one might reject
antiquity's conception of the essence of what is mathematical as
primitive. Yet it has a well founded sense in that for the Greeks
there are simple forms. For them the circle is a simple form to the
extent that it can be produced with an instrument of construction,
and thus it can serve as an element of proof for the mathematical
existence of composite formations. To determine it through an
endless nesting of connected series of line [Polygonzugen]-despite
the affinity of such a conception to pictorial-symbolic intuition-
would have contradicted antiquity's conception of the essence and
uniqueness of mathematical construction.
The fundamental difference between modern mathematics and
that of antiquity consists, as .was already shown, precisely in the
extension of the concept of construction in such a way that it is
transposed to certain operations of thought (algorithms). It is only
through this that an "infinite construction" is acceptable as a
legitimate mode of generation of mathematical formations. Thus for
modern times, the quadratic irrational numbers are no longer the
results of finite geometric constructions, but limit formations of
infinitely converging processes. Moreover, the admission of endless
processes as a justifiable principie of construction of numbers leads
not only to quadratic, but also to other algebraic as well as transcen-
dent irrationalities.
In geometric terms this obviously means that the existence of
mathematical points is ascertained only when a procedure can be
given on the basis of which they are discoverable as limits of number
sequences oras interval nestings. Their specific problem of existente
depends on the "sequence" character of their construction-more
precisely, on the endlessness of the converging processes that ought
to lead to them.
First, let us briefly consider those sequences that are, in the
narrower sense, regular sequences capable of being constructed by a
so-called closed analytical expression.1
54
And yet these are the sale
154. It is to be observed that the concept of the regularity of a sequence is
always relative to specific characteristics. Even sequences whose numbers
are determined by a general number, which thus cannot be presented as
Riemann's Geometry
301
sequences that in an extended sense can be regarded as phenome-
nologically accessible at all. The closedness of the expression allows
such a sequence to be in a certain sense a really "admittable"
sequence. The accessibility of the formation defined by it will here
be guaranteed at least in relation to its foundations of construction,
i.e., if the beginning member of the sequence and thus the initial
interval of nesting are "themselves given," then due to the all-
pervasive regularity of the sequence, they are also constructed, in a
certain sense.
To be sure, there are always only a finite number of discrete
intervals capable of fulfillable postulation, i.e., symbolically intui-
tive fulfillment. It is thereby essential that similar individuals and
discrete postulations be taken one after the other and must be
actually accomplished in time. The factually fulfillable positing
intention of such nesting is thus a sequence of acts whose structure
is finite in principie. The open endlessness of the sequence will not
be attained in it. Yet these acts, which appear necessary in accor-
dance with the idea of nesting, need not be individually fulfillable
acts. The "and so forth," in which the actually positing intention
exhausts itself after a few steps, is not completely indeterminate, due
to the pre-given law of formation of such a succession. Rather, such
a regularity makes it possible to survey the entire succession.
Phenomenologically speaking, such an overview is of a most unique
kind, and in its total structure it differs essentially from all previ-
ously considered "viewing" intentions. In itself it is neither one of
the positing acts of nesting, nor is it a synthetic act comprising its
potential totality. Directed "toward" the open-endless succession,
such an overview is an intention of a second level. More precisely,
the infinite sequence is surveyed with one glance: a singular finite
act of the mathematizing consciousness. It does not intend the
infinite progressively-in accordance with the Hegelian "bad" in-
finity-but, dueto the power of the "and so forth ... ," there is the
categorial fulfillment of the intended as it is ruled and secured
through the law of formation of the sequence. This is the mode in
which the infinite can become a problem for a being who is finite,
but whose consciousness is at the same time such that its variously
merely recurrent, are not simply regular, as if they were determined vis-a-vis
all characteristics. The problematic of the endless is contained in this partial
indeterminateness. Since the endless cannot be surveyed to the end, there is
everywhere a partial indeterminateness of sequences and thus an arithmetic
problem as such.
302 Mathematical Space
stratified structure of acts is capable of constructing and mastering
the infinite in finite thought.
Certainly, in this lies the fact that the infinite can never become a
phenomenon, can never be an object in the strict sense. Thus the
point to which the convergence of nesting "leads" cannot be a really
given and designatable phenomenon. The iterative construction of
sequences-and thus of (sorne) mathematical points-reveals the
fact that these sequences move in a continually open horizon and
remain endless processes. The construction of such points can never
be closed and can only be a constant becoming; in its construction,
the point can be grasped in perpetua! becoming.
This character of becoming in time shows up more impressively
and radically in the so-called freely chosen sequences. After all, not
every point is constructable through a regular or through a determi-
nate recurrent sequence. In the effort to reach "any" chosen point
(but not "all" points!), modern mathematics thought it necessary to
allow sequences in which the succession of the members is not
determined by any pre-given regularity, but rather in which each
member is posited by an act of free choice. Naturally, here only a
finite number of members is factually positable. What is meant by
the "and so forth" in this freely chosen sequence is obviously quite
different in its intentional sense from that of the regular sequence.
While in the latter case the "and so forth" is a kind of direction to
break off further procedure, since in accordance with the rEigularity
nothing "new" can occur, the idea of the freely chosen sequence
contains the notion of a factual progression in infinitum, since its
becoming constantly contains new theses. Yet this idea essentially
abolishes the horizon in which any further postulation should occur.
A mere requirement to posit "something" further is purely formal
and completely abstract-an empty intention that can be fulfilled by
any arbitrary ad hoc postulation; no finite chain of postulations
insufficient. Moreover, there is no positing consciousness having a
capacity for an overview of a higher arder.
The question as to whether such sequences can still be seen as
constructible is essentially a terminological question. If we do not
wish to designate a series of arbitrary postulations without pre-given
rule of procedure as "construction," then the constructive character
explicated up to now must be denied to them. Yet the existence of
points that are to be built through the freely chosen sequences
becomes extremely controversia!- and controversia! not because of
the indefiniteness of the process leading them, but above all due to
the nature and mode of access to them. As was stressed, however, the
Riemann's Geometry
303
question of the existence of the mathematical point can be sensibly
posed only as a question of its access.
155
What the freely chosen sequences offer for the discussion of this
question-and offer not differently from, but more clearly than,
regular sequences-consists above all in clarifying the particularly
mathematical view of infinity in its specific potential character. Just
as a real number is not, but becomes through its generation, so does
its corresponding point. The sequence is not merely a method for
grasping a preexistent point, but rather the manner and mode of
generating it in the first place. And it generates the point only as long
as it is a sequence in process-the point is not just generated within
it. The manner and mode of its generation allows that it will be
generated in infinitum without being generated.
Thus a mathematical point does not exist. At any rate, its exis-
tential claim is not united with any such claim defensible in the
geometric sense; it is not "given," since no construction can actually
"give" it. Even in the regular sequences, a constructible point is, in
a strict sense, "not given." If one insists, its mode of being lies in the
paradox of being a constructively generatable formation only to the
extent that this construction never reaches an end. Not only is all
ontology fooled by it, but it also remains the l'enfant terrible of pure
mathematics "any" point.
The paradox of its existence finally becomes an aporia, however,
"all" points are meant. Mathematics is completely clear about the
risks of this kind of massive postulation. Behind it lies the question
of the mathematical continuum. The problem of what is a genuine
mathematical continuum, and how does it relate to real numbers, is
confronted by mathematics with astounding reservation, and is
enveloped in cautious and not completely clear exposition.
15
B This
must be considered less an expression of perplexity than the positive
155. The mathematical problems of such sequences and the critical
objections raised against them from the side of pure mathematics cannot be
discussed here. The concern here is with what they accomplish for the
mathematical conception of the intuited continuum.
156. H. Weyl, who in many ways has made the continuum into a theme
of mathematical investigation by seeing and clearly formulating the prob-
lematic contained in it, raises the decisive question: "Why do we postulate
under the continuum the concept of real number?" - only to discard it
immediately ("This is not the task here ... ") and to turn for information to
the concept of real number as "the abstract schema of the continuum." See
(1), p. 70-71.
304 Mathematical Space
insight that here the concern is with an "object" of mathematics
whose conception, in contrast to any other mathematical concep-
tion, must receive a strongly modified sense.
If one wanted to see the continuum intended in the sense of
approximation to ever greater exactness the longer the process goes
on, then it would mean that one is imputing a false sense to the
infinite sequence of numbers and respectively to the processes in
consciousness correlated to them. That taken singly these processes
possess precisely definable limits, which are interpretable as points,
does not mean that the continuum consists of points or that it is a
(not enumerable) multitude of points; rather, it allows us to recog-
nize what can occur with the continuum in mathematical treat-
ment-and above all,. what cannot occur. The character of the
sequence as a process running into infinity is mathematically a
precise expression of the possible in infinitum division of the
continuum, which remains a pure possibility. If the point sought in
it were to be actually attained, the continuum would be disrupted.
The condition, that the point never is but always becomes, saves the
continuum. The linguistic expression stating that the continuum can
be "divided" into points has a defensible sense only when one
avoids the paradox of the point and assumes it in the total actuality
of its becoming.
What is most significant for the freely chosen sequence is that this
becoming, as well as its incompletability, becomes directly obvi-
ous.157 They themselves appear only through their becoming, pos-
ited in acts of free choice-and they are only insofar and to the
extent that they are posited. Any member in them is a discrete and
continually new postulation. lt is predetermined and predisposed by
nothing; it is an act of freedom which, in accordance with its idea, is
unlimited in its possibilities. And yet it is an action in time, finite,
factually limited to a finite number of really possible steps in the
framework of duration possessed by a factual subject.
157. Since within the framework of our inquiry the concern is only with
the mathematical conception of the intuitable continuum, we do not
explicate the question any further with respect to what the theory of freely
chosen sequences achieves for the solution of the purely mathematical
problem of the contiunuum. For details see O. Becker (2), pp. 600 ff.
Concluding Observations
The problems treated last were especially concerned with the
constitutive relationship of mathematical objectivity to the mathe-
matizing subject-more clearly: to the subject as a temporal, finite
being. It is an old insight. Aristotle had already recognized the
temporal moment of mathematics in the phenomenon of the se-
quence. Kant stressed the relationship of number to time which for
him was a form of intuition of a finite and indeed specific human
being.
Finally, O. Becker has convincingly treated, in minute phenome-
nological analyses, the modern problematics of pure mathematics
and the decisive role of temporality for the being-characteristic of
mathematical objects. Committed to the transcendental-constitutive
mode of research of Husserl as well asto Heidegger's hermeneutical
phenomenology, his effort is led by the intention of illuminating the
ontological meaning of the mathema, of the mathematical, in an
encompassing sense from the process of "mathematizing," of
mathematikheiesthai, as a mode of the living existence of the human
being.
158
The anthropological founding of mathematics, as seen by
Becker, is not to be understood in a sense of "anthropologism." The
basic structures of mathematics cannot be dissolved into empirically
determined processes of thought, into the psychic acts of the factual
human being. Rather, he intends to show that there are essential
relationships between the sense-structure of the mathematical and
the ontological meaning of the human essence or the idea of a finite
being.
159
The reason why the finitude of the mathematizing subject appears
decisively in this anthropological relationship will be comprehen-
sible directly from the problem of modern mathematics, understood
158. O. Becker (2), pp. 441, 637 ff,; correspondingly (7), especially
pp. 157 ff.
159. O. Becker (3), pp. 279-83.
306 Concluding Observations
as a method for the mastery of infinity: only a finite being can make
sense of the problem of the infinite, only he can want to master it,
and thereby find himself confronted by the abyss of the unsurvey-
able, innumerable, and undecidable.
Our preceding investigation immerses itself in this problematic. It
does so primarily in the sense in which even the geometric domain
ultimately obtains its existence from the mastery of the infinite in
pure mathematics. Just as the latter is related to the being of the
subject in time, geometry too appears to be essentially related to a
being in space: only a spatial and spatially bound being can pose
problems for himself of the kind that lead to geometry as a science of
space.
This relationship is not exhausted in its depth and in its specific
nature by such an analogy alone, as the preceding investigation,
particular! y in the field of the mathematics of space, believed to have
been able to show. Immersed in the specific question of the basis of
being of geometric phenomena, the investigation did not first dis-
cover it in the domain of mathematical syntheses and the accom-
plishment of certain acts of the mathematizing consciousness, but
rather attempted to understand it from preceding achievements of
the subject in his corporeity. In our retrogression to these founda-
tions, we found ourselves placed befare two limits that a phenome-
nologically attained understanding has to respect. On the one side,
there was the doxic-thetic character of consciousness with its
"positing" activities as a contingency not to be inquired behind,
which could not be deduced from the corporeal activities of the
subject. What appeared as uniquely demonstrable was an intercon-
nection of the achievements of the lived body and consciousness
such that in the structure of the latter-contingent in itself-and
specifically in the domain of geometrically constitutive acts, impli-
cations were encountered on every side that could be deciphered
only by retrogression to the functioning of the lived body as such. On
the other side, our phenomenological observations discovered an
impassable limit in the facticity of the corporeal being-more
concretely, in the corporeal constitution of the human. While it was
conceived as a specific concretization of the general principie of
corporeity, in this specification it could not be made further under-
standable, but had to be left in its impenetrable facticity.
If we recall the function that had to be attributed to the lived body
for the constitution of space per se, then the result for us is that the
question of the existence of mathematical spaces-as an ontological
question concerning the mode of their being-will not be answered
Concluding Observations 307
merely in relationship to sorne finite being or other. Rather, it must
be answered more concretely and fundamentally from the mathe-
matizing process of a corporeal human being who is designed in this
and no other way. It is first of all "on the basis" of its corporeity that
this being is a temporal, finite being-it is "on the basis" of this, its
lived body, that it is a spatial being at all. To discover in this the
ultimately founding condition of the geometrical sense of being was
the intent of the preceding investigation.
It has nevertheless served the understanding of this sense only
one-sidedly, because it has not brought to light what occurs through
the multi-stratified constitution of the geometric by the corporeal
human being, and what happens to him himself. The essential
correlative relationship between the full sense-structure of the
geometrical and the total ontological structure of the space-
constituting subject should not be permitted to rigidify into a
statically conceived relationship of arder. Rather, it must be pre-
sented in the dynamics of its reciprocally implicative becoming.
This task was hardly mastered by the means of the phenomenolog-
ically descriptive method. Ultimately, however, we remain indebted
to the subject for illuminating the meaning that must be attributed
not only to the subject's "way" through the diverse spaces, but
finally to his reflection on this way as his own.
Works Cited and Consulted
W. Ahlmann, Zur Analysis der Optischen Vorstellungslebens. Ein
Beitrag zur Blindenpsychologie, Arch. f. d. ges. d. Psychol. Bd. 46,
1924,p. 195-261
Aristotelis Opera, ed. l. Bekker, Berlin, 1831
W. Arnold, Das Raumerlebnis in Naturwissenschaft und Erken-
ntnistheorie, Nrnberg, 1949
H. U. Asemissen, Strukturanalytische Probleme der Wahrnehmung
der Phanomenologie Husserls, Koln, 1957
G. Bachelard, La Potique de l'Espace, Paris, 1958 (German transla-
tion by K. Leonhard, Mnchen 1960
J. J. Bachofen, Urreligion und antike Symbole, Leipzig 1926
O. Becker, (1) Beitrage zur phiinomenologischen Begrndung der
Geometrie und ihrer physikalischen Anwendungen, Jahrb. f. Phil.
u. Phan. Forsch. Bd. 6, 1923, p. 385-560
__ , (2) Mathematische Existenz. Untersuchungen zur Logik und
Ontologie mathematischer Phanomene, Jahrb. f. Phil. und Phan.
Forsch. Bd. 8, 1927, p. 441-809
__ , (3) ber den sogenannten "Anthropologismus" in die Phil-
osophie der Mathematik, .Phil. Anz. Jg, III, 1928, p. 369-387
__ , (4) Grundlagen der Mathematik in geschichtlicher Entwick-
lung, Freiburg, 1954
__ , (5) Das Mathematische Denken in der Antike, Gottingen 1957
__ , (6) Die Archai in der griechischen Mathematik. Einige ergan-
zende Bemerkungen zum Aufzatz von K. v. Fritz, Arch. f. Be-
griffsgesch, Bd. 4, 1959, 210-226
__ , (7) Grosse und Grenze mathematischen Denkens, Freiburg
1960
G. Berkeley, A New Theory of Vision, London 1733
E. W. Beth, ber Lockes "Allgemeines Dreieck," Kantstud. Bd. 48,
1956/57, p. 361-380
309
310 Works Cited and Consulted
L. Binswanger, Das Raumproblem in der Psychopathologie, Ztschr. f.
d. ges. Neurol. u. Ps. 145, 1939; Reprint in: Ausgewahlte Vortrage
und Aufsatze, Bd. 11, Bern, 1955, p. 174-225
M. Blanchot, L'Espace Litteraire, Paris 1955
E. Bleuler, Lehrbuch der Psychiatrie, Zrich 1960
O. F. Bollnow, Das Wesen der Stirnrnungen, Frankfurt 1956
F. J. J. Buytendijk, (1) Allgemeine Theorie der rnenschlichen Haltung
und Bewegung, Berlin 1956
__ , (2) Wege zurn Verstandnis der Tiere, Zrich 1958
__ ,Das Mehschliche. Wege zu seinern Verstandnis, Stuttgart 1959
F. J. J. Buytendijk and H. Plessner, Die Deutung des mirnischen
Ausdrucks. Ein Beitrag zur Lehre van Bewusstsein des anderen
Ichs. Reprint in H. Plessner (3), p. 132-180
R. Carnap, Der Raum, Ergh. d. Kantstud., Berlin 1922
o E. Cassirer, Mythischer, iisthetischer und theoretischer Raum,
Beilageh. z. Ztschr. f. Asth. u. allg. Kunstw. Bd. 25, 1931
J. Cohn, Geschichte des Unend1ichkeitsprob1erns im abendlan-
dischen Denken bis Kant, Darrnstadt 1960
H. Conrad-Martius, Der Raurn, Mnchen 1958
K. Deichrnann, Das Problem des Raumes in der griechischen
Philosophie bis Aristoteles, Halle 1893
R. Descartes, Oeuvres, ed. Adarn Tannery, Paris, 1902 ff.
W. Dubislav, Zur Wissenschaftstheorie der Geometrie, Bl. f. dt. Phil,
Bd. 4, 1930,p. 368-381
K. v. Drckheim, Untersuchungen zum gelebten Raum, Neue
Psychol. Stud. Bd. 6, Heft 4, 1932, 287-473
Euclidis Opera Ornnia, edd. J. L. Heiberg u. H. Menge, Leipzig
1883-1916
G. Ewald, Neurologie unf Psychiatrie, Mnchen-Berlin, 1959
E. Fettweis, Orientierung und Messung in Raum und Zeit bei
Naturvolkern, Stud. Gen, 11. Jg. 1, p. 1-12
E. Fink, Zur ontologischen Frhgeschichte von Raum-Zeit-
Bewegung, Den Haag 1957
P. Finsler, Formale Beweise und die Entscheidbarkeit, Math. Ztschr.
Bd. 25, 1926,p. 676--713
F. Fischer, (1) Raum-Zeit Struktur und Denkstorung in der
Schizophrenie, Ztschr. f. d. ges. Neurol. u. Ps. 124, 1930, p. 243-
258
__ , (2) ber die Wandlungen des Raumes im Aufbau der
schizophrenen Erlebniswelt, Der Nervenartzt, Bd. 7, 1934, p. 84-
86
J. O. Fleckenstein, Die Erweiterung des kosmischen Raumbegriffs in
Works Cited and Consulted
311
der Geschichte der Raummessung, Stud. Gen. 11, Jg, 1, 1958, p.
29-35
K. v. Fritz, (1) Die Archai in der griechischen Mathematik, Arch. f.
Begriffsgesch. Bd. 1, 1955, p. 13-103
__ , (2) Gleichheit, Kongruenz und Ahnlichkeit in der antiken
Mathematik bis auf Euklid, Arch. f. Begriffsgesch. Bd. 4, 1959, p.
7-81
W. Fuchs, Untersuchungen ber das Sehen der Hemianopiker und
Hemiambliopiker, Ztschr. f. Psychol. 1920, I. Teil: Bd. 84, p. 67-
169, 11. Teil: Bd. 86, p. 1 ~ 1 4 3
A. Gehlen, (1) Der Mensch. Seine Natur und Stellung in der Welt,
Bonn 1955
__ , (2) Urmensch und Spiitkultur. Philosophische Ergebnisse und
Aussagen, Bonn 1956 W. Gent, (1) Die Philosophie des Raumes
und der Zeit. Historisch-kritische und analytische Untersuch-
ungen, Bonn 1926
__ (2) Die Raum-Zeit Philosophie des 19. Jahrhunderts, Bonn
1930
G. Gentzen, Die Widerspruchsfreiheit der reinen Zahlentheorie,
Math. Ann. 112, 1936, p. 493-565
K. Gtidel, ber formal unentscheidbare Siitze der "Principia
mathematica" und verwandter Systeme, Monatsh, f. Math. u.
Phys.Bd. 38, 1931,p. 174-198
J. W. v. Goethe, Farbenlehre. Theoretische Schriften, Tbingen 1953
K. Goldstein, ber Zeigen und Greifen, Der Nervenartzt, Bd. 4, 1931,
p. 453-466
K. Goldstein und A. Gelb, (1) Psychologische Analysen
hirnpathologischer Falle auf Grund von Untersuchungen
Hirnverletzter, Iabh., Ztschr. f. d. ges. Neurol.u. Ps. 41, 1918, p.
1-142
__ , ber den Einfluss des vollstiindigen Verlustes des optischen
Vorstellungsvermtigens auf das taktile Erkennen, 11. Abh., aus (1),
Ztschr. f. Psychol, Bd. 83, 1920, p. 1-94
C F. Graumann, Grundlagen einer Phiinomenologie und Psychologie
der Perspektivitiit, Phdn, Psych. Forsch, Bd. 2, Berln 1960
A. A. Grnbaum, Aphasie und Motorik, Ztschr. f. d. ges. Neurol. u.
Ps., 130, 1930, p. 385-412
N. Gnther, Die Struktur des Sehraumes, Stuttgart 1955
A. Gurwitsch, Beitrag zur phiinomenologischen Theorie der
Wahrnehmung, Ztechr. f. Phil. Forsch. Bd. 13, 1959, p. 419-437
N. Hartmann, (1) Zur Grundlegung der Ontologie, Meisenheim 1948
__ , (2) Der Aufbau der realen Welt, Meisenheim 1949
312 Works Cited and Consulted
__ , (3) Philosophie der Natur, Berlin 1950
M. Heidegger, Sein und Zeit, Tbingen 1953
H. Heimsoeth, Der Kampf un den Raum in der Metaphysik der
Neuzeit, Phil. Anz. Jg l, 1925/26, p. 3--42
H. v. Helmholtz, (1) Die Tatsachen in der Wahrnehmung, Berlin
1879
-- (2) Zahlen und Messen erkenntnistheoretisch betrachtet,
Leipzig 1887
E. Hering, Grundzge einer Lehre vom Lichtsinn, Berlin 1920
H. Hermes, Aufzahlbarkeit, Entscheidbarkeit, Berechenbarkeit.
Einfhrung in die Theorie der rekursiven Funktionen, Berlin
Gottingen Heidelberg 1961
D. Hilbert, (1) Axiomatisches Denken, Math. Ann. 79, 1918, p. 405-
415
__ , (2) ber das Unendliche, Math. Ann. 95, 1926, p. 160-190
__ , (3) Neubegrndung der Arithmetik, Abh. d. math. Sem. d.
Universitt Hamburg, Bd. I, 1922, p. 157-163
__ , (4) Grundlagen der Geometrie, mit Revisionen und Ergdn-
zungen van P. Bernays, Stuttgart 1956 ,
D. Hilbert und P. Bernays, Grundlagen der Mathematik, l. Teil Berlin
1934, II, Teil Berlin 1939
O. Holder, (1) Anschauung und Denken in der Geometrie, Leipzig
1900
__ , (2) Die mathematische Methode, Berlin 1924
H. Hoff, Zum Klein Cliffordschen Raumproblem, Math. Ann. 95,
1926,p. 313-339
E. Husserl, Philosophie der Arithmetik, Halle 1891
__ , Logische Untersuchungen, Erster Teil1900
__ , Logische Untersuchungen, Zweiter Teil, Halle 1901
__ , Vorlesungen zur Phanomenologie des inneren Zeitbewuss-
tseins, hg. v. M. Heidegger, Jahrb. f. Phil. u. Phan. Forsch. Bd. 9,
p. 367--496
__ , Formale und Transzendentale Logik, Halle 1929
__ , Erfahrung und Urteil. Untersuchungen zur Geneologie der
Logik, hg. u. red. v. L. Landgrebe, Hamburg 1948
__ , Ges. Werke Husserliana aus dem Husserl Archiv in Louvain,
auf Grund des Nachlasses veroffentlicht unter Leitung van. H. L.
van Breda:
__ , Bd. I, Cartesianische Meditationen und Pariser Vortrige, hg.
v. S. Strasser, Den Haag 1950
__ , Bd. III, Ideen zu einer reinen Phanomenologie und Phano-
menologischen Philosophie; I, hg. v. W. Biemel, Den Haag 1950
Works Cited and Consulted 313
__ , Bd. IV, Ideen ... II, hg. v. W. Biemel, Den Haag 1952
__ , Bd. V, Ideen ... III, hg. v. W. Biemel, Den Haag 1952
__ , Bd. VI, Die Krisis der europaischen Wissenschaften und die
transzendentale Phanomenologie, hg. v. W. Biemel, Den Haag
1954
__ , Bd. VIII, Erste Philosophie, Teil II, hg. v. R. Bohm, Den Haag
1959
__ , Die Frage nach dem Ursprung der Geometrie als
intentionalhistorisches Problem, mit einem Vorwort van E. Fink,
Rev. Int. d. Phil. 1. Jg No. 2, 15. 1. 1939, p. 207-225
E.R. Jaensch, (1} Zur Analyse der Gesichtswahrnehmungen, Ztschr.
f. Psych. u. Physiol. d. Sinnesorgane, Ergbd 4, 1909
__ , (2} ber die Wahrnehmung des Raumes, Ztschr. f. Psych. u.
Physiol. d. Sinnesorgane, Ergbd 6, 1911
M. Jammer, Concepts of Space, New York 1954 (Germ. transl. by P.
Wilpert: Das Problem des Raumes, Darmstadt 1960}
H. Jantzen, ber den kunstgeschichtlichen Raumbegriff, Sitz.-Ber
d. Bayer. Akad. d. Wiss., Phil. Hist. Abt., Jg. 1938, H. 5
K. Jaspers, (1} Zur Analyse der Trugwahrnehmungen. Leibhaftigkeit
und Realita.tsurteil, Ztschr. f. d. ges. Neurol. u. Ps., 1911. p. 460-
535
__ , (2) Die Phanomenologische Forschungsrichtung in der
Psychopathologie, Ztschr. f. d. ges. Neurol. u. Ps, 1912, p. 391-
408
__ , Allgemeine Psychopathalogie, Berlin 1946
E. Kahn, ber Innen und Aussesn, Mtsschr. f. Psychiatr. u. Neurol.
1955, 1. Teil: Bd. 129, p. 171-177; 2, Teil: Bd. 130, p. 375-380
W. Kaiser, (1} Das sprachliche Kunstwerk, Bern 1959
__ , (2} Entstehung und Krise des modernen Romans, Stuttgart
1955
I. Kant, Ges. Schriften, hg. v. d. K%umonigkl-Preuss. Akad. d.
Wissensch., Berlin 1902
D. Katz, (1} Die Erscheinungsweise der Farben und ihre Beeinflus-
sung durch die individuelle Erfahrung, Ztschr. f. Psychol. u.
Physiol. d. Sinnesorgane, Ergbd. 7, 1911
__ , (2} Der Aufbau der Tastwelt, Ztschr. f. Psychol. Ergbd 11, 1925
__ , (3} The World of Colour, London 1935
W. Killing, (1} ber die Clifford Klein'schen Raumformen, Math.
Ann. 39, 1981,p. 257-278
__ , (2} Einfhrung in die Grundlagen der Geometrie, Paderborn
1893
K.P. Kisker, Der Erlebniswandel der Schizophrenen. Ein
314 Works Cited and Consulted
psychopathologischer Beitrag zur Psychonomie schizophrener
Grundsituationen, Monogr. a. d. Gesamtgeb. d. Neur. u. Ps. Heft
89, 1960
L. Klages, Ausdrucksbewegung und Gestaltungskraft, Leipzig
1923
F. Klein, (1) ber die sogenannte Nicht-Euklidische Geometrie,
Math. Ann. I. Teil Bd 4, 1871, p. 573-625; Il. Teil Bd 6, 1873, p.
112-145
__ , (2) Vergleichende Betrachtungen ber neuere geometrische
Forschungen, Erlangen 1872 and Math. Ann. Bd 43, 1893, p. 63-
100
__ , (3) Zur Nicht-Euklidischen Geometrie, Math. Ann. Bd 37.
1890,p. 544-572
K. Kleist, ber Apraxie, Mtsschr. f. Psych. u. Neurol. 19, 1906, p.
269-290
]. Konig, Sein und Denken. Studien im Grenzgebiet von Logik,
Ontologie und Sprachphilosophie, Halle 1937
A. Koyre, From the Closed World to the Infinite Universe, Baltimore,
1957
H. Kronfeld, ber neuere patopsychisch-phii.nomenologische
Arbeiten, Ztrlbl. f. g. ges. Neurol. u. Ps. 28, 1922, p. 441---459
H. Khn, Die Kunst Alteuropas, Stuttgart 1954
].H. Lambert, Theorie der Parallelismus 1766, Erstabdruck im C. F.
Hindenburgs Mag. f. r. u. angew. Math. f. 1786
L. Landgrebe, Prinzipien der Lehre vom Empfinden, Ztschr. f.
Philos. Forsch. Bd. VIII, 1954, p. 195-209
H. Lassen, (1) Beitrage zur Phanomenologie und Psychologie der
Raumanschauung, Wrzburg 1939
__ , (2) Subjektiver Anschauungsraum und objektiver Gegen-
standsraum in der Kantischen Philosophie, Ztschr. f. dt.
Kulturphil. Bd. 6, 1940, p. 15---41
K. Lewin, Der Richtungsbegriff in der Psychologie, Psychol. Forsch.
Bd 19, 1934, p. 249-299
W. Lietzmann, Anschauliche Topologie, Mnchen 1955
P.F. Linke (1) Phii.nomenologie und Experiment in der Frage der
Bewegungsauffassung, Jahrb. f. Phil. u. Phan. Forsch. Bd 2, 1926,
p. 1-17
__ , (2) Grundfragen der Wahrnehmungslehre, Mnchen 1929
G. Lintowski, Typische Einstellung bei Wahrnehmungsleistungen,
Ztscr. f. Psychol. u. Physiol. d. Sinnesorgane, 120, 1931, p. 126-
188
Works Cited and Consulted
315
J. Locke, An Essay concerning human Understanding, ed. by A. St.
John, London 1902
F. Lobell, (1) ber die geodatischen Linien der Clifford-Klein'schen
Flachen, Math. Ztschr. 30, 1929, p. 572-607
__ , (2) Ein Satz ber die eindeutigen Bewegungen der Clifford-
Klein'schen Flachen in sich, Jahrb. f. r. u. angew. Math. 162, 1930,
p. 114-124
__ , (3) Zur Frage der geodatischen Linien in den offenen Clifford-
Klein'schen Flachen mit positiver Charakteristik, Jahrb. f. r. u.
angew. Math. 162, 1930, p. 125-131
P. Lorenzen, Einfhrung in die operative Logik und Mathematik,
Berlin, Gottingen, Heidelberg 1955
G. Marcel, Etre et Avoir, Paris 1935
G. Martin, Neuzeit und Gegenwart in der Entwicklung des
mathematischen Denkens, Kantstudien, Bd 45, 1953/54, p. 155-
165
F. Mayer-Hillebrand, ber die scheinbare Grosse der Sehdinge,
Ztschr. f. SinnesphysioJ. 61, 1930/31, p. 267-324
M. Merleau-Ponty, (1) PhenomenoJogie de la Perception, Paris 1945
__ , (2) La Structure du Comportement, Paris 1953
W. Metzger, Gesetze des Sehens, Frankfurt 1953
S. Monat-Grundland, Gibt es einen Tastraum? Ztscr. f. PhysioJ. u.
Psycol. d. Sinnesorgane 115, 1930, p. 209-271
W. Morgenthaler, Die Abbau der Raumdarstellungen bei
Geisteskranken, Beilageh. z. Ztscr. f. Asth. u. allg. Kunstwiss. Bd.
25, 1931
L. Nelson, Bemerkungen ber die nichteuklidische Geometrie und
den Ursprung der mathematischen Gewissheit, Abh. d. Fries-
'schen Schule, Neue F. 1906, p. 386-392
O. Neugebauer, Comptes Rendus du Congres InternationaJ des
Mathematiciens, Bd. l. Oslo 1957
M. Palgyi, WeJtmechanik, Ges. Werke, Bd. III, Leipzig 1925
r. Pauli, Der Aufbau der Tastwelt, Arch. f. d. ges. Psychol. Bd. 56, H
2, 1926, p. 253-280
B. Petermann, ber die Bedeutung der Auffassungsbedingungen fr
die Tiefen und Raumwahrnehmung, Arch. f. d. ges. Psychol. Bd.
46, 1924, p. 351--416
A. Portmann, Biologische Fragmente zu einer Lehre vom Menschen,
Basel1951
H. Plessner, (1) Die Einheit der Sinne. GrundJegung einer Asthesi-
oJogie des Geistes, Bonn 1923
316 Works Cited and Consulted
__ , (2) Die Stufen des Organischen und der Mensch. Einleitung in
die philosophische Anthropologie, Berlin 1928
__ , (3) Zwischen Philosophie und Gesellschaft, Bern 1953
A. Podlech, Der Leib als Weise des In-der-Welt-Seins, Bonn 1956
Proklos, In Euclidem, ed. G Friedlein, Leipzig 1873; dt. bersetzung
v. L. Schi:inberger, hg. v. M. Steck, Halle 1945
H. Read, Eingeborenenmalereien, New York 1954
K. Reidemeister, (1) Anschauung als Erkenntnisquelle, Ztschr. f.
Phil. Forsch. Bd. 1, 1946, p. 197-210
__ , (2) Das exakte Denken der Griechen. Beitrdge zum Deutung
van Euklid, Plato, Aristoteles, Hamburg 1949
__ , (3) Raum und Zahl, Berlin, Gi:ittingen, Heidelberg 1957
H. Rein und M. Schneider, Physiologie des Menschen, Berlin,
Gi:ittingen, Heidelberg 1960
E. Rothacker, Probleme der Kulturanthropologie, Bonn 1948
F.S. Rothschild, ber rechts und links. Eine erscheinungswis-
senschaftliche Untersuchung, Ztschr. f. d. ges. Neurol. u. Ps. 124,
1930,p. 451-511
B. Riemann, ber die Hypothesen, welche der Geometrie zugrunde
liegen, Abh. d. Konigl. Ges. der Wissensch. zu Gottingen 1867,
unveranderter fotomech. Nachdr. Darmstadt 1959
F. Sander, Die Entwicklung der Raumtheorien in der zweiten Halfte
des 19. Jahrhunderts, Diss. Halle 1931
J.P. Sartre, L'Erte et le Neant, Paris 1947
M. Scheler, (1) Die Wissensformen und die Gesellschaft, Leipzig
1926
__ , (2) Wesen und Formen der Sympathie, Bonn 1931
__ , (3) Idealismus-Realismus, Phil. Anz. 2. Jg, Heft 3, 1927, p.
255-324
P. Schilder, (1) Das Korperschema. Ein Beitrag zur Lehre vom
Bewusstsein des eigenen Korpers, Berlin 1923
__ , (2) Fingeragnosie, Fingerpraxie, Fingeraphasie, Der Nerven-
arzt, Bd. 4, 1931, p. 625-629
F. v. Schiller, ber Anmut und Wrde, W. d. Inselausg. Bd. 5,
1938
C. Schneider, ber Sinnestrug, Ztschr. f. d. ges. Neurol. u. Ps., 1931;
l. Teil: Bd. 131, p. 719-813, 11. Teil: Bd. 137, p. 458-521
K. Schneider, Die phanomenologische Richtung in der Psychiatrie,
Phil. Anz. 1. Jg. 2. Hbbd, 1925/26, p. 382-404
H. Scholz, Die Axiomatik der Alten, Bl. f. dt. Phil. IV, 1930; Abdruck
in: Mathesis Universalis, Basel1961, p. 27-44
Warks Cited and Cansulted 317
P. Schroder, Gefhle und Stimmungen, in Wissen und Scheidewege
van Leben und Geist, Festschr. f. L. Klages, Leipzig 1932, p. 201-
222
A.v. Senden, Die Raumauffassung der Blindgebarenen var und nach
der Operatian, Diss. Kiel1931
Th. Skalem, ber die Nichtcharakterisierbarkeit der Zahlenreihe
mittels endlicher oder abzii.hlbar unendlich vieler Aussagen mit
auschliesslich Zahlenvariablen, Fund Math, 23, 1934, p. 150-
161
A. Speiser, Die mathematische Denkweise, Basel1952
P. Stii.ckel und F. Engel, Die Thearie der Parallellinien van Euklid
bis auf Gauss, Leipzig 1895
A.D. Steele, ber die Rolle von Zirkel und Lineal in der griechischen
Mathematik, Quel. u. Stud. z. Gesch. d. Math., Phys. u. Astr. Abt.
B. 3, 1936,p. 287-369
E. Stier, Untersuchung ber die Linkshandigkeit, Jena 1911
S. Strasser, Das Gemt. Grundgedanken zu einer phanamenologi-
schen Philasaphie und Thearie des menschlichen Gefhlslebens,
Utrecht, Antwerpen, Amsterdam 1956
E. Straus, (1) Die Formen der Rii.umlichen, ihre Bedeutung fr die
Motorik und die Wahrnehmung, Der Nervenarzt, Bd. 3, 1930,
Abdruck in: Psychalagie der menschlichen Welt, Berlin,
Gottingen, Heidelberg 1960, p. 141-178
__ , (2) Vam Sinn der Sinne, Berlin 1956
E. Stroker, Die Perspektive in der bildenden Kunst. Versuch einer
philosophischen Deutung, Jb. f. Asth. u. allg. Kunstwiss. Bd. IV,
1958/59, p. 140-231
C. Stumpf, ber den psychalagischen Ursprung der Raumvarstel-
lung, Leipzig 1873
E.v. Sydow, Die Kunst der Naturvolker und der Varzeit, Berlin 1923
H. Tellenbach, Die Rii.umlichkeit der Melancholischen, Der
Nervenarzt, Bd. 27, 1956, p. 12-18; 289-298
H. Tischner, Kunst und Sdsee, Hamburg 1954
J.v. Uexkll, (1) Theoretische Bialagie, Berlin 1928
__ , (2) Streifzge durch die Umwelten von Tieren und Menschen,
Hamburg 1956
P. Valery, L'Anne et la Danse, Paris 1924
W. Voss, Subjektive und Objektive Aufbauelemente in den
Zeichnungen Blinder, Hamburg 1931
V.v. Weizsii.cker, (1) Der Gestaltkreis. Thearie der Einheit van
Wahrnehmen und Bewegen, Stuttgart 1950
318 Works Cited and Consulted
__ , (2) Zwischen Medizin und Philosophie, Gottingen 1957
__ , (3) Natur und Geist, Gottingen 1957
H. Werner, Raum und Zeit in den Urformen der Knste, Beilageh. z.
Ztschr. f. Asth. u. allg. Kunstwiss. Bd. 25, 1931
H. Weyl, (1) Das Kontinuum, Leipzig 1918, 1921
__ , (2) Mathematische Analyse des Raumproblems, Berlin 1923
__ , (3) Raum, Zeit, Materie, Berlin 1919, 1923
__ , (4) Philosophie der Mathematik und Naturwissenschaft,
Mnchen Berlin 1927
__ , (5) Die Stufen des Unendlichen, Jena 1931
W. Wieland, Die Aristotelische Physik. Untersuchungen ber die
Grundlegung der Naturwissenschaft und die sprachlichen
Bedingungen der Prinzipienforschung bei Aristoteles, Gottingen
1962
J. Wittmann, ber Raum, Zeit und Wirklichkeit, Arch. f. d. ges.
Psychol. Bd. 47, 1924, p. 428-511
H.G. Zeuthen, Die geometrische Construction als "Existenzbeweis"
in der antiken Geometrie, Math. Ann. Bd. 47, 1896, p. 222-228
J. Zutt, Rechts-Links St6rung. Konstruktive Apraxie und reine
Agraphie. Ein Beitrag zur Pathologie der Handlung, Mtsschr. f.
Psychiart. u. Neurol. 82, 1932, p. 253-305; 355-395
Register
W. Ahlmann, p. 128.
64
Aristotle, p. 157, 161 ff.
W. Arnold, p. 96.
36
H.U. Asemissen, p. 100.
41
J.J. Bachofen, p. 34.
14
O. Becker, p. 11, 184,
92
193, 211,
109
220,
116
246,1
26
261 ff., 266,
1
38
297,
151
305.
G. Berkeley, p. 95, 98, 101 f., 123.
E.W. Beth, p. 197.
101
O.F. Bollnow, p. 20.
3
F.J.J. Buytendijk, p. 16,
2
31,
9
33, 35.
19
R. Carnap, p. 182 f.
E. Cassirer, p. 35.
40
J. Cohn, p. 162.
81
H. Conrad-Martius, p. 7,
2
162.
81
R. Descartes, p. 201, 203, 243.
K. v. Drckheim, p. 15,
1
28.
2
Euclid, p. 211,
109
213, 217 ff., 227, 231 ff., 240, 265 ff., 270, 286,
298 ff.
F. Fettweis, p. 74.
29
E. Fink, p. 7,
2
175.
87
P. Finsler, p. 219.11
5
K. v. Fritz, p. 211,
109
267.
138
W. Fuchs, p. 133.
68
A. Gehlen, p. 34.
14
W. Gent, 7.
2
G. Gentzen, p. 220.
117
K. Godel, p. 220.
J.W.v. Goethe, p. 22.
K. Goldstein and A. Gelb, p. 78,
32
127 ff.
319
320
C.F. Graumann, p. 83.
33
A.A. Grnbaum, p. 71.27
N. Gnther, p. 96.
36
A. Gurwitsch, p. 107.
46
Register
N. Hartmann, p. 166,
84
171 f., 256.
131
M. Heidegger, p. 2, 4, 8 f., 16,
2
41 f., 51 f., 53,
21
385.
H. Heimsoeth, p. 7.
2
H. Hermes, p. 219.
115
D. Hilbert, p. 200, 207 ff., 213, 220, 227 f., 269.
O. Holder, p. 197,
101
213.11
2
H. Hopf, p. 259.
132
E. Husserl, p. 4 f., 8, 11 f., 38,
17
42, 83,
33
88,
34
98 f., 104 ff., 123, 150,
166,
84
171 f., 175, 180, 184, 187,
93
191 ff., 196,
99
197, 202,212,219
ff., 245,
125
305.
E.R. Jaensch, p. 96.
36
M. Jammer, p. 7,
2
162.
81
W. Kaiser, p. 21.
4
l. Kant, p. 1, 2, 105, 126, 165 f.
D. Katz, p. 124
58
, 126.
W. Killing, p. 260,
134
261,
135
L. Klages, p. 22.
F. Klein, p. 259, 265, 275 ff.
J. Konig, p. 22.
5
A. Koyre, p. 7.
2
J.H. Lambert, p. 268 f., 274 f.
L. Landgrebe, p. 107.
48
H. Lassen, p. 94,
35
131, 153 ff.
K. Lewin, p. 72.
J. Locke, p. 123, 188,
94
196.
F. Lobell, p. 259.
132
G. Martin, p. 220.11
8
M. Merleau-Ponty, p. 2, 16,
2
58, 99 ff., 158 ff.
S. Monat Grundland, p. 128.
63
L. Nelson, p. 213.11
2
O. Neugebauer, p. 211.
109
M. Palgyi, p. 37, 42, 126.
R. Pauli, p. 126.
60
A. Podlech, p. 17.
2
A. Portmann, p. 31.
9
H. Plessner, p. 16,
2
58.
23
Proklos, p. 267 f., 275.
K. Reidemeister, p. 202,
104
209,
108
212.
E. Rothacker, p. 16.
2
F.S. Rothschild, p. 65.2
5
Register
B. Riemann, p. 269, 274, 285 ff.
Sacceri, p. 268.
F. Sander, p. 96.
36
J.P. Sartre, p. 17,
2
58.
23
W. Schapp, p. 124.
58
M. Scheler, p. 16,
1
44, 116 ff., 158 ff., 166,
84
181 ff.
P. Schilder, p. 65,
25
71,
27
78,
32
127.
F.v. Schiller, p. 33.
H. Scholz, p. 211.
109
P. Schroder, p. 20.
3
A.v. Senden, p. 128.
64
Th. Skolem, p. 220.U
8
A. Speiser, p. 211.
109
P. Stackel and F. Engel, p. 272.
142
A.D. Steele, p. 217.U
4
S. Strasser, p. 20.
3
E. Straus, p. 21,
4
29, 36.
E. Stroker, p. 22,
4
125.
59
C. Stumpf, p. 101,
42
153.
H. Tellenbach, p. 44.
19
J.v. Uexkll, p. 31.
9
P. Valery, p. 35.
13
W. Voss, p. 135.
69
V.v. Weizsacker, p. 31.
9
H. Weyl, p. 242,
123
288, 303.
156
W. Wieland, p. 158.
79
J. Wittmann, p. 128.
64
H.G. Zeuthen, p. 217.
114
J. Zutt, p. 78.
32
321

Вам также может понравиться