Sei sulla pagina 1di 61

Husserl Studies 20: 207–267, 2004.


C 2004 Kluwer Academic Publishers. Printed in the Netherlands.

Questions Regarding Husserlian Geometry


and Phenomenology. A Study of the Concept
of Manifold and Spatial Perception∗

LUCIANO BOI
Ecole des Hautes Etudes en Sciences Sociales, Centre de Mathématiques,
54 boulevard Raspail, 75006 Paris, France

Denn diese Art strenger Raumordnung im ganzen beruht darauf, dass jedes
lokale Geschehen von dem anderen, man möchte fast sagen ‘dynamisch
weiss’ (W. Köhler, Die physischen Gestalten in Ruhe und in stationären
Zustand, 1923).

It belongs to the sense of this unity [the unity-consciousness], which we


call the thing, to be unity in a manifold of appearances, in a continuity of
appearances of a determinate, ideal, infinite type. And the thing is given
in every actually streaming piece of this continuity. The thing is also given
in every appearance of such a piece, and in each one the consciousness of
the identity of the fulfillment is alive, even if it is a changed appearance
. . . But it pertains to the essence of this givenness to leave open infinitely
many possibilities of new givenness in a determinate way as motivated
possibilities (E. Husserl, Ding und Raum, 1907).

A thing can exist for me only in the indissoluble unity of sensation and
intuition, by the fact that a continuum of quality covers a [spatial-temporal]
continuum of extension (H. Weyl, Mind and Nature, 1934).

Any theory of spatial perception aims at explaining the relationship between


perceptual, physical and geometrical spaces. Husserl contributed to this in
an essential way, particularly by showing the type of proto-geometry that is
involved in the structuring of phenomenal world or in the constitution of the
“spatial thing.” Here, we aim at showing that the concept of manifold is central
to understand the nature of these relationships. In fact, the sensory space, we
could say the ‘natural environment’ of our perception, constitutes a type of
amorphous topological continuum that, as well as being a primitive datum of
our intuition, presents a qualitative structure of primary importance. Gener-
ally, it is possible to show that any fundamental property that characterizes the
208
mathematical concept of manifold finds its equivalent in the spatial percep-
tion’s field. In the same line of thought, we bring to attention the role played
by the mathematical concept of group in the structuring of the movements of
three-dimensional objects in the external world as apprehended in perception.
But Husserl’s attempt only reached its goal partially. The main reason being,
we believe, that this analysis stop at a descriptive phenomenological level and
that, therefore, he does not take seriously into account the possibility of mak-
ing mathematically intelligible the phenomenal field of spatial perception.
There exists a fundamental opposition that seems to us at the heart of the phe-
nomenological project: that between real space (the surrounding Euclidean
space) and space as an abstract manifold (ideal space). Unlike Riemann and
Weyl, Husserl thinks that this last kind of space has no other meaning than a
formal one and he rejects that it could get an intrinsic or real meaning. We
will criticize this point of view and object the following: (i) that, since it is
essentially the real world that is geometric and not geometry that is real, even
phenomenal space can be endowed with interesting geometrical properties, as
it seems to be the case; and (ii) that nothing prevents us to think that real space
supports some geometrical “beings” that we will characterize as “manifolds”
that will serve as models for substratum spaces in which we define a set of
sense qualities, the dynamic evolution of which we can follow in time.

1. Introductory Remarks on Spatial Perception

Perception is the nodal point of any theory of reality that aspires to intelligibil-
ity. Perception is the “site” of encounter between the world of physical objects,
the world of images founded on those objects and of the representations of real
phenomena that have been developed thereon. The expression “nodal point”
here is not intended solely as a metaphor. Rather, if the concept is taken in its
precise mathematical sense, wherein a knot is defined, in the simplest case,
as an imbedding1 (a continuous map) of a curve S 1 in R 3 , it provides us with
a direct analogy for understanding the content of perception, say, as a kind of
“imbedding” of a set of physical objects pertaining to the visual field into the
global space of perceptive structures. (Let us add that though it may remain
necessary in the quasi-generality of cases for an object to be seen, this condi-
tion in and of itself is an insufficient one for cognitive perception. The object
must also insert itself into a determined perceptual structure in accordance
with a particular spatial and temporal order.) In either case, the operation does
not amount to the identical sum of its two elements, but to some essentially
new thing. Both cases involve a process that is dynamic rather than static, and
just as the knot may produce a new space out of a certain space, shaping it
in a certain way, a coherent series of perceptions leads to a representation of
phenomena that is only indirectly linked to physical objects as they exist in
the exterior world, and which, incidentally, is only partially conditioned by the
209
constitution of our sensory systems. We can nevertheless say with certainty
that the very process of perception would be unintelligible if we were to make
complete abstraction of the nature of the physical causes and the physiological
organization of our perceptual system. This is an important point that we shall
try to develop later.
Let us now consider the process of perception as such. All perception
is perception of spatial forms located in space. These forms underlie certain
properties pertaining to the space they inhabit, but the nature of our perception
is such that only a small number of these are detected. For example, continuous
evolutions – the slow transformations that generally result in the metamor-
phosis of an organism, that is to say a change in the nature and structure of a
form – are only perceived with considerable difficulty under normal circum-
stances. We are more likely to perceive sudden changes, especially when these
entail one or more qualitative discontinuities. This introduces the matter of the
different levels of perception; we will summarize them in the following way:
(i) the pre-spatial level (the different organized series of sense impressions);
(ii) the two-dimensional spatial level: i.e., those of the surfaces’ objects
imbedded in our ambient space (the usual Euclidean space) and directly
perceived by our senses;
(iii) the objective spatial level (three-dimensional); this is the level at which
spatial forms, proper, are constituted, as a consequence of our ability to
discriminate a certain number of important parameters (endowed with
a certain permanence) and to disregard those that appear contingent or
accidental;
(iv) the symbolic and/or scientific level that extends perception beyond its
natural limits and seems fundamental to the understanding of invisible
structures (such as microscopic ones) or phenomena that occur at a very
large scale to be perceptible (such as the behavior of astral bodies).
To cite an example, electromagnetic interactions are so weak that, on the
cosmic as well as on the human scale, a glance cast on the physical system
proves ineffectual. Conceptual observation makes a fundamental intervention
in such an instance. The eyes of the body are essentially replaced by the “eyes
of the mind” of which Galileo spoke, in an effort to see not that which can
presently be seen but that which should or could be seen (see on this subject
Hanson, 1958). To return to the above example, borrowed from physics, this
type of vision should be capable of “seeing” in the phase-space of mechanics,
in the space of elementary events of probability theory, in the four-dimensional
curved space–time of general relativity, in complex projective space of infinite
dimensions of quantum theory. To understand what is visible to “real eyes,”
we need only know that a section or a silhouette of a universe of infinite
dimensions is projected on the retina. Naturally, such a point of view carries
with it the consequence that only through the invisible can the visible be
210
apprehended, thus transforming the world of phenomena into one of pure
appearances (Plato’s theory). Such an approach is primarily interested in not
limiting the scientific knowledge of world’s structures to directly or indirectly
observable entities, and extending it in consideration of a very large universe of
conceptual or fictitious “observables,” or more precisely “non-observables”
such as phase function-psi, virtual photons, the color of a quark, and the
difference between identical particles of quantum mechanics.
It seems, however, equally (if not more) reasonable – and this is the per-
spective that we are trying to develop here – to consider the following: first,
the phenomena that we ordinarily observe do not conceal any less mystery
or meaning with regard to the structure and real properties of the universe.
Second, the phenomena we perceive in a very rich variety of apparent forms
are really traces through which, by means of appropriate modeling or formal-
ization, it becomes eventually possible to reconstruct the complete structure
of objects. This amounts to saying that what we perceive does not pertain to
the order of pure appearance because, by virtue of appearing, it is an aspect
of the nature of being that is manifesting itself.
Of course, what appears to us cannot single-handedly constitute even the
content of our representations or scientific knowledge of the world, even
though these may follow from it. Symbolic and/or scientific perception is also
important, for the precise reason that it strips perception of all its dependent
causal relations with a certain physical reality, which is nonetheless at its
origin, in order to construct “polysemic” images of this same reality. Several
different systems are implicated in perception, but they are not all reducible
to an underlying physical–chemical system; in fact, each one of them exhibits
a certain autonomy. Be that as it may, they are intimately correlated and work
together to form a global space of perception; there exists a set of relations
between these systems that act in accordance with a functional hierarchic
structure.
Let us return for a moment to the subject of the existence of different levels
of perception, in order to pinpoint an element which strikes us as important.
The phenomenon of perception presents a certain constant character:2 for
example, the invariance of the perceived dimension of an object, independently
of the distance at which it is seen. When an object displaces itself towards the
eyes, the size of its image on the retina increases, even though the object is
not incorrectly perceived as growing in volume. Color perception is another
example of the constancy of the phenomenon of perception, as when the color
of an object is perceived independently of the kind of light that is cast on it.
And finally, there is the invariance of the perceived position of an object as
when there is a voluntary movement of the head or eyes. Now, the ability to
abstract the constancy of color, position and size of an object from the images
that form on the retina constitutes, in fact, an initial level of the process of
perception. A much more highly developed level would consist in the ability
211
of abstracting the constancy of form. When an object is seen under variable
circumstances, its perceived form maintains itself constant, even if totally
distinct types of sensory receptors are stimulated at diverse moments. This
variability of the particular aspects of an object does not even necessarily
reach consciousness. It is the form of the object that is abstract, regardless of
the specific part of the sensory system that provides the initial information
about the object. It is this capacity to abstract the form of an object out of the
(necessarily discrete) multitude of information bits that allows us to speak of
the perception of forms (Gestalt), that is to say of the object as a whole. In
fact, this capacity is one of the fundamental requisites for the formation of
a category of objects, and it is doubtless one of the earliest concepts to be
formed in the course of the development of the human intellect.
Consequently, perception is inseparable from the problem of form recog-
nition. In principle, this idea can be expressed in a purely mathematical for-
mulation. The question at hand is essentially one of defining the equivalence
class of a system of objects of the same nature, or of recognizing the “spatial
forms” that pertain to the same equivalence class.3 If P is a set of equivalent
forms – that is to say closed domains of Euclidean space E – well defined
abstractly, then how do we know if a closed domain  belongs or not to this set
P. Two forms F1 and F2 will be said homomorphic (topologically equivalent)
if there exists a homeomorphism φ of E on E so that F2 = φ(F1 ). Gener-
ally speaking, we will say that two spatial forms M and N are topologically
equivalent if we can move from one to the other through a continuous map
(deformation) φ : M → N . We shall say that the objects of the external world
(the closed forms of surrounding Euclidean space) that fulfill this definition,
present the same form type. From this point of view, form recognition does not
require the intervention of conscious activity. It may occur, as is the case for
certain animal species (for example, fledglings), due to an innate mechanism
that releases a reaction to certain key stimuli that are markedly pregnant and
characterized by a very simple relationship. But, as adulthood progresses, par-
ticularly in the human species, the recognition of perceptual forms is generally
accompanied by a conscious activity, and even by an increasingly complex
mental activity as the forms become less and less simple (for example, in
what other way can a spatial multi-connected form be recognized?). It is at
this point that the mental phenomenon of perceptive constancy intervenes.
There are, in addition, a number of physical hypotheses as well as con-
straints imposed by the perceptive apparatus to be taken into consideration
with respect to the recognition of forms. Among these, our inability to
perceive beyond a certain limit must be explained by spatial-temporal
objective constraints linked to certain physical phenomena such as the type
of physiological organization of our sense organs. In this respect, it is widely
known that the retina operates as a receptor of wavelengths emanating from
an electromagnetic field that responds to a rather narrow interval. In fact, a
212
photographic plate is far superior in this regard. The fact that our eye reacts
within the limits of wavelengths on the order of 6.000 Å is, in no doubt,
linked to the fact that natural light, by virtue of which living organisms on the
earth see objects, and which is basically the spectral decomposition of solar
rays, has its maximum intensity in the neighborhood of this middle range.
This proves, among other things, that sensory perception is limited because it
depends, to some degree, on certain physical facts (as we have just seen), and
to some degree on the constitution of our perceptive system itself. Sensory
perception can therefore only reconstruct a relative and (approximate) value
of reality, and can never yield an absolute or certain value.
A profound understanding of how the morphological level of reality is con-
stituted and of the precise relation it shares with perception is of fundamental
interest. We can already say that the emergence of forms obeys two different
processes that can be concomitant: the formation of continua by merging and
the fragmentation of the surrounding perceptual field (for example, the visual
field) by qualitative discontinuities. In reference to this problem, we must men-
tion the question of the relationship between spatial extension and sense qual-
ities (or qualia). One must understand how these two elements are modified
as a function of each other. (These points shall be pursued in more depth in the
pages that follow.) Let us introduce three aspects that pertain to this question:

1. Given an electromagnetic field endowed with certain physical properties,


all sense qualities (such as color) emerge out of their own substratum space
to constitute themselves as objects of perception within the visual field by
way of a continuous series, or, more simply a continuum; the understanding
here is that all continua can be seen as forms.
2. The jump from one perceptive form to another or the transition from a
continuum to another may be brought about by at least one qualitative
discontinuity (or singularity) occurring in the substratum space of all sense
qualities.
3. All forms or continua are connected wholes, global autonomous units. The
merging of their connected components constitutes them and a new form
can be generated from the product of two disjoint connected components
that, in turn, can also be seen as forms.

We know that this concept plays a major role in topology. The fundamental
relationship of these continua will therefore be that of the part to the whole
and not that of the element to the set. Incidentally, as the whole is more than the
sum of its parts, it cannot be obtained from a simple juxtaposition of its parts.
Nor can we reproduce a parent form from a recomposition of its elements after
they have been divided, because such a method involves the destruction of an
essential unity. These continua form global units, and the principles that rule
them can exist independently of the nature of the parts that constitute them.
213
The visual system plays a privileged role in perception in general and in the
recognition of forms in particular. The reasons for this are to be found first and
foremost in the following facts: (i) the anatomic and physiological structure
of the visual organ; and (ii) its functional organization. Although it is retinal
vision, in the first instance central and peripheral, that enables the detection
of forms, selective recognition of all forms requires a central synthetic vision.
Furthermore, it is important to note that vision possesses a certain underly-
ing geometry. The visual system constitutes a two-dimensional manifold that
possesses a projective rather than a metric structure. More specifically, there
exists a certain isomorphism between projective geometry and the geometry
of vision, which is performed by the retina. Let us add here that the behavior
of light and, more specifically, the geometric and optic laws, which it obeys,
are responsible, in a large measure, for the way in which we visually perceive
the forms of objects.4
A fundamental epistemological problem arises in all theories of spatial
perception: is the difference between psycho-physiological space (perceptual,
representative) and geometric space (scientific, ideal), a difference in essence
or solely in degree? If we opt for the latter, we are accepting the idea, according
to which geometric space is basically obtained through an abstract exfolia-
tion of the properties that characterize psycho-physiological space, thereby
allowing both to contribute to the formation of global perceptual space. Then
again, we may suppose, in this case, that ideal geometric space is in reality the
virtual prolongation of representative space, now freed of all kinds of limits
and constraints that attach it to real (present) perception. After all, we con-
sider that perceptual space, like geometric space, has its source in the whole
of biological mechanisms developed by the species and the individual in view
of their relationship to the environment and their mastery thereof. But it is
nonetheless clear that such a position winds up overlooking the fundamental
characteristic of all geometric space as the result of the active construction of
a pure concept, a product of thought that justifies itself independently of all
relations to the primitive oculomotor space of primates or even to the more
complex mental space of our species. We need only think, in this regard, of the
myriad metric spaces, topological and functional invented by mathematics in
the last century with an exclusively theoretical goal. This shedding of mental
and physical evolutionary constraints has led to a goodly number of abstract
mathematical spaces (that in the quasi-generality of cases are remote from
any physical intuition) characterized, for example, by the use of the meter to
determine the distances between objects. Used since antiquity as a practical
instrument to estimate the distances between rigid bodies on a human scale,
it would later become a standard of ideal length to measure imperceptible
physical distances, among other things, before being replaced by other units
of abstract measurement or by mathematical concepts such as the function of
distance (metric).
214
Finally, we underline another important epistemological point. Any sensory
perception is actually from the outset integrated to an inherently representa-
tional and eventually symbolic process that in due course culminates in an
activity which is properly conceptual and scientific. According to certain psy-
chologists and philosophers, such a distinction would in fact lack any founda-
tion. These thinkers exclude the very possibility of a sensory perception, as it
is our system of concepts or our language that since the beginning has directed
and determined what we incorrectly designate as “the contents of perception.”
Without entering here into this debate,5 it seems to us, all things considered,
that a strict idealist position in this case, as in others, encounters a major dif-
ficulty which we shall sum up as follows. It needs to be acknowledged that
thought cannot reproduce itself indefinitely without thereby becoming an im-
mense tautology, unless we accept the idea that it must rest on certain contents
of understanding which we shall call intuitions, judgements, concepts, etc.,
and that these last must in their turn be founded on mental acts such as the
imagination, attention and reflection. We are consequently obliged to accept
that these in turn emerge, at least in part, from various perceptive contents
(not to be confused with sense impressions) and are exercised with a view to
the constitution of true objects of perception. In order to avoid this difficulty,
it is necessary to postulate a philosophical principle based on a fundamental
critical realism: there exists a correlation between our thought contents and
the real structures of the world. This correlation rests on a certain number of
physical and mathematical invariants that characterize the real world, but that
are also “internalized” by thought. It is thanks to these, in fact, that we are
capable of recognizing in the realm of diverse and apparently incoherent phe-
nomena a certain geometric and morphological organization and emerging
bundles of meaning that refer to it.

2. The Concept of Manifold and the Constitutive Levels of Space: From


Riemann to Weyl

Hermann Weyl has singled out the understanding of the relationship between
intuitive or sensory (psycho-physiological), geometric (conceptual, ideal)
and physical (abstract although physically determined) spaces, respectively,
as one of the most fundamental outstanding scientific and epistemological
questions.6 Taking up Husserl’s expression, he adds that these three types of
space form the constitutive levels (konstitutive Stufen) of the general con-
cept of space. In point of fact, Weyl is directly responsible for the demon-
stration of two important insights. He has shown that geometric space (his
prototype R 3 , or R n , or even, any open set of R n ) is different from surround-
ing Euclidean space E (das extensive Medium der Außenwelt, in its proper
sense) in that it is an ideal mathematical concept. He has also shown that
215
physical space is a mode of actualization (or a deformation) of geometric
space; we can, in fact, in principle conceive of a number of possible defor-
mations. To this end, we could think, along with Poincaré, that representative
space, too, is a deformed image of geometric space that primarily obeys the
laws of optics, perspective, or even physiology. But such a position seems to
us unsatisfactory, even if it highlights a significant aspect of representative
space.
In contradistinction to Kant and in agreement with Riemann, Weyl main-
tains that it is not geometric space as such, and especially not its axioms, that
are an a priori given; instead, it is the space of our immediate experience (in
other terms, surrounding space) constituting a sort of amorphous topological
continuum, that is a primitive given of our intuition. We might even say, in
fact, that we uncover its qualitative nature as soon as we relate it to the per-
ception of objects in the external world. Geometric space, on the other hand,
can be obtained, according to Weyl, by a process which consists of supplying
that amorphous continuum with a certain number of essential mathematical
structures that mutually implicate each other: metrical, differential, topolog-
ical, etc. It was Riemann who first developed an entirely new conception of
mathematical space on which differential geometry and topology, as well as
a large part of modern physics, rests. But given its much more general range,
especially for the understanding of relations between the concept of manifold
and our phenomenal as well as physical space, it appears useful to review its
essential ideas below.
First, the concept of manifold (Mannigfaltigkeit) seems rather funda-
mental.7 A manifold can be intrinsically defined as any part of space en-
dowed with a proper geometric structure. Any surface will then be conceived
as a two-dimensional manifold equipped with certain fundamental geometric
and topological properties, independently of its specific relationship to the
surrounding Euclidean space, hence as a mathematical object in itself. Sim-
ilarly, a line will be considered as a one-dimensional manifold, and space as
a three-dimensional manifold. We can, therefore, safely extend our concep-
tion of manifolds into higher-number and even infinite dimensions and treat
them as autonomous mathematical objects. But not all surfaces or manifolds
exhibit the same mathematical properties. The study of these properties has
led to a new fundamental distinction between metric (local) properties and
topological (global) properties. As a consequence, it became clear for the
first time that manifolds that are equivalent in terms of their metric prop-
erties (e.g., a plane and a cylinder) can be altogether different with respect
to their topological properties. Accordingly, these mathematical objects now
present themselves to our mind in a classification that is based on their in-
trinsic properties and their respective structures. In this way, we have grasped
the extremely important idea that two manifolds that are metrically, or topo-
logically, equivalent constitute what would henceforth be known as a same
216
spatial form. The surfaces or n-dimensional manifolds that can be continu-
ally deformed one into the other while their metric relations remain invariant
are accordingly called isometric. Those, on the other hand, that can be devel-
oped one over the other through a continuous global deformation are called
homeomorphic (topologically equivalent).
The above implicitly includes three other fundamental ideas introduced by
Riemann. It would be useful to highlight their importance here, especially as
this will lead to a better understanding of certain remarks made by Husserl on
the nature of space, and more particularly, of spatial perception.
(i) All manifold may be endowed with several metric structures (determi-
nations); from this point of view, Euclidean space may be treated as one
among other possible determinations.
(ii) Manifolds can be distinguished with respect to whether they are con-
tinuous or discrete; the first are endowed with a metric (differentiable)
structure; the second possess an essentially algebraic structure. The fore-
going also supports the distinction between the relations of extension
or region (Ausdehnungs-oder Gebietsverhältnisse) and metric relations
(Maßverhältnisse). An important, but not the only, consequence of such
a distinction is that we can conceive of different metric relations for the
same relations of extension.
(iii) All manifolds may be provided with many kinds of structures that are
at once metric, differentiable, topological, algebraic, etc.; this is what is
principally responsible for their mathematical interest.
One must also include the following idea which is related to point (ii),
and which, incidentally, is applicable not only in physics but also in
certain domains of the natural and human sciences, particularly spatial
perception.
(iv) This idea concerns the distinction between the “unboundedness” (Unbe-
grenztheit) and the “infinite” (Unendlichkeit). The first is a topological
property; the second a metric property. In other words, we can very well
conceive of a manifold as being finite and limitless at one and the same
time. One need only think of the example of the sphere S 2 . One can go
around it an unlimited number of times, but as topologically it is closed
and all its points are mutually identical, we can only trace finite move-
ments because we can not leave the space it encloses without breaking
its boundary and, no matter where we start our movement we will return
exactly to our point of departure. In short, we are speaking of a model of
a finite space. A passage from Riemann appears particularly enlightening
on this point:
That space is an unbounded three-fold manifold, is an assumption which is
developed by every conception of the outer world; according to which every
instant the region of real perception is completed and the possible positions
217
of a sought object are constructed, and which by these applications is for
ever confirming itself. The unboundedness of space possesses in this way
a greater empirical certainty than any external experience. But its infinite
extent by no means follows from this . . .8

In this regard, let us recall that the idea that the laws characterizing optic
space were the same as those pertaining to a finite two-dimensional space
had already been formulated towards the end of the 18th century. The Scot-
tish philosopher Thomas Reid, among others, while examining the geometric
properties of the visual field that he believed to be spherical, recognized
important analogies between the geometry of the sphere and the geometric
structure that characterizes vision. From this he was led to make the dis-
tinction between tactile geometry, which is Euclidean, and the geometry of
vision which follows other than Euclidean laws. Today, the theories available
to us on this subject are more precise (see below for more remarks on this
matter).
We shall later see how the ideas summed up in points (i) through (iv) find
their equivalents in the domain of the visual field, and, more generally, in that
of spatial perception. In fact, they play an important role in the analyses made
by Husserl to characterize the structure of the objective space of perception.
Let us limit ourselves here to a few remarks:

(i) It is not yet ascertained if the sensory systems forming the general process
of perception are equivalent from the metrical point of view. Put another
way, it is not certain that “visual space,” “motor space” or “auditory
space,” conceived as subspaces of the global space of perception (for
this notion, see infra), possess the same metric structure; it also cannot
be excluded that each of them may allow sensitive variations of their
metric parameters, depending on the environmental characteristics and
functional circumstances in which they act.
(ii) We can distinguish between continuous and discontinuous fields of per-
ception. The first can be characterized (in a general manner) as sets of
sense qualities, that, while allowing one or several continuous variations,
remain preserved with respect to their essential form. The visual field,
for example, is a continuum in the sense that the series of sense impres-
sions that accompany the sight of an object, first of an aspect then of its
integrity form a continuous series (ein Kontinuum der Wahrnehmung, as
C. Stumpf has written).9 Similarly, the domain of colors, geometrically
illustrated by the disc giving the full spectrum together with laws of com-
position, is a clear example of what we understand here as continuous;
we can in effect obtain any new color of the spectrum based on a series
of variations of particular primary colors. A tonal system is also a con-
tinuum for certain parameters of sound can be made to vary, even if the
218
principal tone is kept constant. Alternatively, a discontinuous perceptual
field is characterized by the fact that its fundamental unity is broken at
some place. Thus, if we completely or suddenly change the sound in-
tensity of a tonal system, it will cease to be perceived as a continuum.
There is, however, an even more fundamental aspect to this question.
Spatial perception is essentially the perception of forms, that is to say,
of objects “filled” with sense qualities. From this point of view, a form
is therefore a phenomenon which appears. For an object to be perceived
as a form, however, it must detach itself from the rest; it must, in other
words, set itself apart from surrounding space by its boundary. The emer-
gence of forms or the passage of one form into another is thus necessarily
accompanied, in perception, by a qualitative discontinuity. In summary,
continua of perception are constituted as a result of a process of gra-
dation and of internal compositions, that is to say of merging; whereas
discontinuous fields are constituted by separation and fragmentation, by
leaps.
(iii) Each perceptual field, it would seem, is capable of realizing several
integrated functions at once; yet, this would be seemingly difficult to
explain without allowing for a differentiation of the physiological and
anatomic structures of respective sensory apparatus. For example, vision,
which constitutes the dominant sense in the apprehension of space, can
treat different spatial information (be it of proprioceptive origin or of
external provenance) pertaining to the distances between objects, their
location, the reconstruction of paths to move from one object to the
next (orientation) as well as to approximation of the relative positions of
bodies in space. There, in regard to space, the predominance of vision
over other sensory modalities is as much qualitative (topological) as it
is quantitative (metrical).
(iv) The differentiation between the “unboundedness” and the “infinity”
character of a manifold finds an immediate analogy in the distinction
between a closed but unbounded space, on one hand, and an infinite
space of possible perceptions on the other. Hence, the vision of a spatial
body corresponding to a succession of visual impressions, each of which
exposes one facet, accompanied by a consecutive series of movements
(for example, rotations) whether of an object around our eye or our body
around the object, is in itself a closed process. Evidently, we need only
change the orientation of the object relative to our body, or reverse the
order in which we perform rotations around the object that remains fixed,
to obtain another sequence of visual perceptions of the same object that
would be equivalent to the first. In fact, it is possible “to go around” in
different ways. Each of these series always corresponds uniquely to the
totality of the actual perception of the object and constitutes the identity
of a particular spatial body. In Husserl’s words:
219
The spatiality of the individual body, as well as space itself, “present them-
selves.” The body is always given in an aspect (view). A closed system of
possible aspects constitutes the complete appearance (Vollapparenz), i.e.,
the complete appearance of the surface of the body, as this appearance
is constituted without my moving (or being moved) from my place. This
appearance of the surface is itself, however, only one appearance in a man-
ifold, and this manifold consists of the universality of identical complete
appearances . . . (quoted from E. Husserl, Thing and Space [thereafter TS]
p. 292).

Following René Thom, one can state this problem in a general way that
invokes the mathematical notion of form. A spatial form F separates itself
in principle from its background, generally Euclidean space E. In this sense,
the notion of form always presupposes a qualitative discontinuity, namely,
one which separates the points of space inside of the form from those on the
outside of the form. Furthermore, we usually accept that a form is topologically
a closed set of surrounding space; any bounded series of points of F admit a
limit point within F. If E is space R 3 , let S be the set of oriented directions of
R 3 ; then for any s ∈ S let C(s) be the apparent contour of form F seen from
infinity according to direction s. Then F (the form) is defined by the totality
of its apparent contours: G F = ∪s CsF (F).
But, to return to Husserl, if we consider this same object, not as it appears
to us at the moment, but as it could appear to us at another given moment, or
in other physical circumstances (for example, of luminance) or under other
spatial relations, such as proximity and distance, or extension and depth,
a new group of rotations, etc.; then we are dealing with an infinite (three-
dimensional) space. In other words, this new space emerges when image
changes of a new kind are related to groups of movements or groups of
kinesthetic data. Naturally, what we have here is an eminently spatial infinity.
With respect to its temporal unfolding, while perception refers only to the
present, this present always carries an infinite past behind it – the retentional
intentionality – and an open horizon before it.

3. Spatial Perception and the Concept of Group; Physiological and


Mathematical Foundations of Space

Perception is necessarily linked to the organism’s capacity for sizing up, in-
terpreting and eventually carrying out certain spatial processes in order to
achieve mastery of its immediate environment. Whether its objectives are bi-
ological, intellectual, aesthetic, or other, the recognition of forms constitutes
doubtless the fundamental operation of this process. Still, the knowledge of
a larger environment that is effectively beyond our reach and our capacity of
observation and immediate action demands in addition a more scientific kind
of understanding that depends on ad infinitum repetition and ideal extension
220
of these very spatial operations. In fact, these are our only available means
for the passage from present to future perceptions, especially to those that are
inaccessible; in a similar way, we attempt to deduce invisible forms from the
recognition of apparent phenomena.
A single example can elucidate this point from the mathematical point of
view. The congruent maps of Euclidean space form a group of transformations
E + , known as the group of Euclidean movements. The foregoing, in fact,
allows us to define in abstracto three-dimensional objects (volumes) as parts
of space that can be carried in any other part of the same space by virtue of a
transformation S of E + . This fact suggests the following interpretation: the
group E + of the congruent maps (of surrounding space) expresses an intrinsic
structure of space itself which it impresses on all spatial objects. If we gener-
alize further, we can easily show that the fundamental concepts of a geometry
can be established at the same time as the group  of automorphisms that
should be regarded as given a priori. In this way, the notion of equivalence can
be defined explicitly. Suppose that F is a subgroup of E + belonging to ; then
the following proposition can be stated: two figures (although, we can easily
substitute sufficiently symmetrical three-dimensional objects here) are equiv-
alent relative to F if one of them can be brought to coincide with the other
through a transformation S of F. We shall furthermore add that such a relation
of equivalence has an objective meaning if, and only if, the F-equivalent
figures can, by virtue of any transformation S of  be brought to coincide
with other F-equivalent figures. In this way, transformations fulfilling to such
a property and belonging to F form a group that is an invariant subgroup
of .
It is well known that these abstract notions of mathematical group theory
introduced notably by E. Cartan and H. Weyl early in the 20th century are
identical with those that rule the behavior of a great number of natural phe-
nomena (in crystallography, in chemistry and certainly in physics), and that
they have enabled us to explain and classify their properties and fundamental
laws. In this regard, we would now like to discuss two groups of problems.
In the first group, we will touch on certain physiological and perceptive as-
pects that are linked to the concept of space and associated notions such as
movement. In the second group, we shall deal from a more mathematical per-
spective with the question of the relationship between the concept of group
and spatial perception.

1. Nineteenth century attempts to study the physiological origins of the con-


cepts of distance and movement led sometimes to remarkable theoretical
and experimental results. First, let us recall the laws stated by E.H. Weber
and G. Fechner. The first, known as the “law of differential threshold”
stipulates that for every kind of sensation there is a constant ratio between
the intensity of the initial stimulus and the minimal variation that it must
221
undergo for the difference to be felt (Weber’s threshold). The second law
provides the exact formula for the relationship between (psychic) sensa-
tions and (physical) stimuli; it states precisely that sensation varies as the
logarithm of the stimulus. Among other hypotheses, it was stated that the
primary elements that participate in the formation of tactile-motor space
are not the concepts of a straight line and a plane, but rather those of a
circle and a sphere. In this way, it was ascertained that tactile-motor space,
particularly as generated by the special organ of the hand, corresponds
to metric space, which as we know, fundamentally rests on the notion of
distance. The fact that metric space is unbounded and that straight line
movement is equally unbounded with respect to it must be explained by
the fact that the organ of tactile sensation can always be moved so as to
produce a succession of equal and indefinitely extended lengths. Accord-
ing to the philosopher and mathematician Federigo Enriques, when certain
data of psycho-physiology are interpreted mathematically, they lead to an
interesting conclusion; namely, that the three groups of representations that
correspond to the theory of continuity (analysis situs), to metric (and dif-
ferential) geometry and to projective geometry can be associated with the
three groups of principal sensations: general tactile-muscular sensations,
sensations of the special tactile organ (the hand) and sight, respectively.
However, there was at the time a broad consensus to deny the foundation
of geometric space on physiological space, for the latter is in fact devoid
of the properties of homogeneity and isotropy that characterize the former.
Furthermore, physiological space was regarded as not continuous, in the
sense that it is not mathematically infinite.
In this respect, the examples of Helmholtz, Mach and Poincaré are suf-
ficiently paradigmatic, even though one must also recognize the important
differences that separate their basic scientific hypotheses as well as their
philosophical conclusions. Poincaré, especially, pursued a line of reason-
ing which, such as it is, could strike us today as partly lacking justification,
and which stated that:

Our sensations differ from one another qualitatively, and they can there-
fore have no common measure, no more than can the gramme and the
metre. Even is we compare only the sensations furnished by the same
nerve-fibre, considerable effort of the mind is required to recognise that
the sensation of today is of the same kind as the sensation of yesterday,
but greater or smaller; in other words, to classify sensations according
to their character, and then to arrange those of the same kind in a sort
of scale, according to their intensity. Such a classification cannot be
accomplished without the active intervention of the mind, and it is the
object of this intervention to refer our sensations to a sort of rubric or
category pre-existing in us. [. . .] This category becomes necessary to us
only for comparing our sensations, for reasoning upon our sensations.
It is therefore rather a form of our understanding. [. . .] Sensible space
222
has nothing in common with geometrical space. [. . .] Seeing that our
representations are simply the reproductions of our sensations, therefore
we cannot image geometrical space. We cannot represent to ourselves
objects in geometrical space, but can merely reason upon them as if
they existed in that space. [. . .] Geometrical space, therefore, cannot
serve as a category for our representations. It is not a form of our
sensibility. It can serve us only in our reasonings. It is a form of our
understanding.10

Certainly, no one would reasonably assume that different sensations con-


sidered in isolation would by themselves give us the notion of geometric
space, nor that geometric space itself is fundamentally equivalent by na-
ture with what we commonly call sensible space. Incidentally, in the above
passage, Poincaré is responsible for pointing to a fact which is of increas-
ing importance today, namely that a certain mental activity (which he calls
mind) intervenes at the earliest phase of perception and that sense impres-
sions become meaningful to us insofar as they form (after having been
treated and “purified” by our perceptive systems) groups of sensible qual-
ities by means of and in which we recognize the objects of the external
world. In this sense, we can say that the qualities of color, sound, etc. are the
categories by virtue of which we achieve a primary structuring of an oth-
erwise chaotic and incoherent sensible world, and from which we proceed
to its “mise en forme.” But by the same token, it seems both unfounded and
naive to speak of crude sensations devoid of law or structure, as though
it were not these very qualitative differences that provide an index of a
certain organization as well as a criterion enabling us to recognize their
various different regimes, and to think that they are lacking in all spatial
character. Still, we readily accept that sensory fields, particularly the visual
field, to which all visual sensations and perceptions refer, exhibit a precise
and extremely interesting (geometric) spatial structure as Husserl, among
others, has clearly demonstrated (see infra).
Generally speaking, in a representation of space, we are dealing with a
whole of organized specific sensations (light, color and shape in the case of
sight). In this regard, we shall recall here the position expressed by William
James and Karl Hering which is at variance with Poincaré’s. Hering is re-
sponsible for the development of an innate theory of spatial perception and
vision in particular, in opposition to Helmholtz’s empiricist theory. James
asserted that each sensation has an implicit spatial character. In reference to
Hering, he later suggested that each sensation, through the mediation of the
stimulated part, corresponds to a certain position, and as in general more
than one and even several parts are present, we can in this sense speak of
a volume of sensations. Ernst Mach also analyzed the differences between
physiological space and metric (geometric) space, arriving at conclusions
that, in the present context, need to be fully appreciated.11 He understands
223
the physiological foundation of spatial perception in the following manner.
First, he introduces the distinction between “sense impressions” (Sin-
nesempfindungen), that correspond to the sensations communicated by an
elementary organ and which depend in part on the sort of stimulation, and
the “organ sensations” (Organempfindungen), that occur when the activity
of the organ is primarily determined by its individual quality; this quality
remains the same with each new stimulation and varies only from organ to
organ, and it can be considered as identical to spatial sensation. One can
reasonably suppose that the more extensive the ontogenetic affinity of the
sense organs is, the more constant (stable) the spatial sensation will be, and
conversely, the less extensive it is the more variable it will become. Organ
sensation (or spatial sensation), even if induced by the stimulation of an el-
ementary organ (such as sight), can remain the same when the same organ
or organ complex is stimulated and the same network of organs is activated,
even if the nature of the stimulus has changed in the meantime. One can say
then, following Mach, that physiological space is a system of “graduated
organ sensations” (abgestuften Organempfindungen), which when aroused
by variable sense impressions will form a “permanent register” (“bleiben-
des” Register). This space is, in a sense, an anatomical–physiological cat-
egory, to borrow Poincaré’s phrase, within which these variable sense im-
pressions are placed. We can, in other words, represent it as the total ex-
tension of the scale of one type of sense qualities, for example, of sound.
In effect, due to what we call the “timbre” of a compounded sound, the ear
is completely capable of discerning the different proportions in which the
principal tone and its harmonics mix, that is to say, it is sensitive to the dif-
ferent sounds of a chord.12 Given what we have just outlined, it is now pos-
sible to pose the problem of the understanding of the relationship between
phylogeny and ontogeny and their respective influence on the development
of spatial perception in more general terms. But this is obviously outside
the scope of the present analysis (see on this subject Boi’s book, 2004).
The ordered character of perceptions, independently of their specific
content, had already been stressed by J.F. Herbart by means of the notion of
serial form (Reihenform),13 almost in the sense of the mathematical concept
of manifold which Riemann would develop some 30 years later. Herbart has
highlighted the fact that a series of perceptions maintains itself not only by
mirroring the material of the individual element of sense impressions, but
also as a series, a precisely ordered sequence of memory. Similarly, words
are not founded solely on articulated sounds, but on precise sequences of
articulated sounds; in this form they are remembered and understood with-
out being, for a minute, confused with the various anagrams which they
can supply. It is a fact that once the global apperception of a given series
of perceptions is completed, all the partial (sensible) representations that
comprise it will henceforth form an intensive unity; had they followed a
224
different sequence, these perceptions would form an identical unity. In this
sense, spatial perception contains the representation of something sequen-
tial (and continuous) as its attribute. We will notice here a certain resem-
blance to the analyses that Husserl developed later. Another aspect, this
time highlighted by Mach, addresses with greater precision the relationship
between the notion of movement and of physiological space. According to
Mach, spatial sensations essentially serve to guide movements.14 Once the
primary experiences regarding the (physical) body, the distances, etc., are
acquired, they become necessary, as “intentional objects,” to the mastery of
surrounding space. Consequently, a being that is immobile or incapable of
changing his orientation will be totally unable to construct a representation
of Euclidean space; his space will always be anisotropic and limited. Yet,
the fact that the body as a whole can move at will and orient itself means,
according to Mach, that the same movements can be carried out everywhere
and in all directions, and that everywhere and in all directions space is of
an identical constitution and can be represented as unbounded and infinite.
But we shall dispute this as precisely not being the case, because
the surrounding space that is within our body’s reach is limited, and
incidentally, in more ways than one, and is fundamentally finite. By virtue
of its constitution, our body experiences further objective constraints, at
once of a biological and mental nature, that ensure that its movements
are far from being free in all directions and everywhere. Nor is the
body capable of many orientations that can generally be impressed on a
solid ideal (geometric) body. A certain lack of symmetries in the human
body is well known; furthermore, we are only capable of completing
certain movements and not others, and those movements which we are
capable of possess a privileged directional axis. Nor are our movements
interchangeable, and their elements cannot be exchanged at will. In
short, it would be difficult to recognize in the movements that under
normal conditions characterize our body the same movements that form
a geometric group of transformations, even the relatively simple one of
Euclidean displacements. We are forced to conclude, therefore, that our
spatial experiences only approximate that of Euclidean space without ever
completely achieving it. It is also indubitable that sensible space lacks
certain characteristically mathematical properties – such as continuity –
that we ascribe to geometric space. As Clifford has clearly pointed out:
. . . I could not possibly see either the surface of a thing, or a motion,
as continuous; for the sensitive portion of my retina, which receives
impressions, is not it self a continuous surface, but consists of an
enormously large but still finite number of nerve filaments distributed
in a sort of network.15
On this basis, Clifford’s line of thinking proceeds towards saying, more
or less, that our spatial sensations are formed of two truly distinct parts.
225
There is a “message” that reaches us from somewhere, in one way or
another (depending on the type of stimulus and the kind of organ being
stimulated),

but this message is not all that we apparently see and hear and feel. In
every sensation there is, besides the actual message, something that we
imagine and add to the message. This is sometimes expressed by saying
that there is a part which comes from the external world and a part
which is supplied by the mind. But however we express it, the fact to be
remembered is that not the whole of a [spatial] sensation is immediate
experience [. . .]; but that this experience is supplemented by something
else which is not in it [by a supplement of information] (Ibid).

According to Clifford, the notion of geometric space would be formed on


the basis of a similar kind of process. We have to complete the properties
and distance relations of solid bodies in reality, in such a way that they
answer to fulfill a code of rules, some called common notions, and some
called definitions, and some called postulates. It is in this way that we later
imagine these same properties and relations in two- and three-dimensional
space. But what do these rules exactly consist of? In answer to this, Clifford
writes:

For example, I sometimes imagine that I see two lines in a position


which I call parallel. Parallelism is impossible on the curved picture of
my retina; so this is part of the “fulfillment”. Now whenever I imagine
that I see a quadrilateral figure whose opposite sides are also equal.
This equality is also a part of the fulfillment, and relates to possible
perceptions other than the one immediately present (Ibid).

2. Let us now proceed to the second question that we had set ourselves. Only
recently we have come to understand the role played by the concept of
group in the structuring of the movement of three-dimensional objects of
the external world as they are apprehended by perception. Inspired by a
seminal work of 1938 (see references) in which Ernst Cassirer highlighted
the relationship existing between the concept of group and the theory
of perception, a number of psychologists and scientists (among others,
Hoffman, Foster, Shepard) sought to prove both theoretically and experi-
mentally that the preferred motions of our body and of our mental activity
reflect an “internalization” of certain principles that govern the trans-
formations of objects in the external world. These principles are of two
kinds:

(a) The laws of physics – and notably the principle of least action,
according to which, in the absence of external forces, a rigid body
follows a straight line path, which is to say that it carries out a rotation
226
around its center of gravity, which in turn moves according to a
rectilinear translation.
(b) Principles of kinematic geometry – and notably Chasles’s theorem,
according to which for any pair of positions in which an asymmetrical
object is located16 there exists a unique spatial axis such that the
object can be carried from one position to any other by a motion that
results from a translation and a rotation around that axis.

We can, however, equally consider a family of intermediate cases and show


that each of them can be formulated abstractly in terms of geodesic paths
(which are transformed parallel to themselves) defined on a six-dimensional
manifold (or a group of transformations), one that includes all possible po-
sitions of objects in space. In effect, as each position of an object in three-
dimensional space can be identified by a unique space transformation that
carries this object from one given position to another, this six-dimensional
manifold can be considered equivalently either as a manifold of all possible
positions of the object, or as a manifold of all rigid displacements that can
be undergone by that same object. In other words, it can be considered as a
manifold of all possible applications of a three-dimensional Euclidean space
E 3 on itself. In this way, the displacement of an object is understood, not as
a motion of the object relative to a fixed space, but rather as the transport of
the object (immobile in itself ) by a rigid motion of the entire space.
It is known, on the other hand, that one of the most important problems
faced by all theories of perception, and by visual perception in particular, is
the problem of the constant character of perceptual phenomena. This problem
is presented as the invariance of the perception of the object’s position, when
it is subject to voluntary movements of the eyes or the head or both simul-
taneously. Several theoretical hypotheses and experimental facts suggest that
these two types of movement be, in fact, geometrically and functionally cor-
related in accordance with a central cognitive schema. Phenomenologically,
the foregoing can be described as follows: as an individual moves his gaze or
turns his head, the image of objects in the surrounding visual field is displaced
on the retina. This movement is not perceived, however, as a movement of the
objects; these are perceived as stationary. Such a compensation of the percep-
tive system for the movement of the image is sometimes so involuntary as to
not be consciously perceived as a movement of the head or eyes. Not only
does this prove that certain properties of invariance that we tend to ascribe to
the objects themselves are, in fact, the result of an important compensatory
activity of a proprioceptive nature and that in the course of this process the
response of each sensory apparatus is de facto integrated into the activity of
the global perceptual system, but also that perception is animated by a move-
ment that is essentially intentional, and that consciousness is consciousness
of space and time, even thought this one cannot be reduced to these.17
227
Let us add one more remark with respect to the problem of the constant
character of perceptual phenomena. We can expose this problem as one that is
much more general in nature, given its relevance to the human mind’s capacity
for filtering and transforming stimuli, especially visual ones, that have their
origin in external objects, in order to obtain information about the observed
and therefore perceived object that is “objective,” that is to say, independent
of any individual. Without in the least wanting to refute the possibility of
several other explanations for such a phenomenon or the participation of
several processes of a physiological and mental nature therein, we would like,
nonetheless, to suggest a purely mathematical interpretation. If we choose
to understand that the constant character of our perceptions is nothing but
the possibility of admitting the identity of the perceived object, we can then
define this identity with the help of the concept of groups in the following
manner. Let θ be the object and G the principal group of all its possible
movements in space. We then call the automorphisms i of G, the isomorphic
mappings of this group onto itself. Every automorphism  of G is a one-to-
one mapping of G onto itself, hence a transformation of the set G. In this
way, two automorphisms can be multiplied and the resulting product gives
a transformation of the group G which is also an automorphism of G. It is
clear, moreover, that the identity application is an automorphism, and that a
transformation inverse to an automorphism is yet again a group. The set of
all the automorphisms of a group Gis also an automorphism. Therefore, the
set of all automorphisms i of a group G forms a group. According to what
precedes, we can say that an object is constant or that it conserves its identity
if all the automorphisms that we may act on and all the operations applied
to them will again form an isomorphic group. Abstractly, the identity of the
object is said to be maximal if its group of automorphisms comprehends
the largest number of reciprocal isomorphic subgroups, and conversely, it
is said to be minimal if the same group contains the smallest number of
subgroups.
Apart from movements that are reflexive or unconsciously induced, we
must consider the very large class of the changing positions of objects that
are compensated for by a voluntary movement of our body. These movements
can be characterized by virtue of their allowing the object’s return to its initial
position after it has been displaced in any way. Such movements, further-
more, allow the indefinite repetition of the same movement, so that despite
the fact that the object undergoes several types of movements in space, it
always seems invariant. (Of course, we are neglecting here the possibility that
the whole of space can also undergo certain movements, such as a change of
orientation, for example.) These objects form the category of solid bodies.
From this point of view, our body together with its different sense organs
constitutes an “ideal” system of reference to which all the objects located in
the surrounding space as well as their relative mutual positions are referred,
228
within the limits, naturally, of a certain number of subjective and objective
constraints. As a matter of fact, as we have seen earlier, perception would
be inconceivable for completely immobile beings. Yet, this second type of
movements exhibits a certain internal composition, a structure obeying pre-
cise laws: of homogeneity, isotropy, symmetry, etc. These movements then
form a group which is the group of the displacements of objects in the space
of perception. This group can, in turn, provide for the existence of subgroups
that are required to be isomorphic to the “principal” group. For the time being,
let us merely remark that several great mathematicians and physicists, includ-
ing Mach and Poincaré, have attempted to explain the genesis of geometric
space on the basis of the notion of the group of displacements acting on the
set of oculo-kinesthetic impressions, only to conclude that geometric space
does not have that much in common with perceptual space, after all. This
type of conclusion remains unclear, however, given that it asserts the com-
pletely formal character of Euclidean geometry, while recognizing Euclidean
space as a homogeneous space of the group of displacements. Furthermore,
the indispensable role played by the set of muscular sensations in the process
of constitution of perceptive space must also be emphasized. In effect, these
sensations function as a veritable sense organ (the so-called sixth sense). Not
only do they help in the activity of other sense organs, such as touch, vi-
sion and kinesthetic and vestibular systems, but they are also the source of
proprioceptive sensibility, which can play a constitutive role in the represen-
tation of the body. Husserl has described very well certain phenomena that
must be considered in relation with the foregoing, in, among other works
Phänomenologische Untersuchungen zur Konstitution, written between 1912
and 1928, and Ding und Raum of 1907.

4. The Intrinsic Structuring of the Phenomenal World According to


Husserl and the Opposition Between Descriptive and Exact Sciences

Husserl is above all interested in analyzing the intrinsic structuring that is


inherent to the phenomenal world. This structuring, however, can not be re-
duced to its physical (material) component, active element though the latter
may be, nor can it be seen as solely a product of the subject or of subjective
consciousness. For it is precisely this structuring of the phenomenal world
that constitutes the milieu out of which significant structures and mental acts
forms. But at the same time, it does not essentially differ from the constitu-
tion of the spatial world through the medium of the perceptual process. From
Husserl’s writing, we can infer without much effort an assertion that the phe-
nomenal world (and its manifestations) is situated at the interface between the
properly physical world (that of Cartesian extension) and the properly mental
world of representations and conceptual thought. The world of phenomena,
which is also that of perceived objects and apparent forms, while being a
229
tributary of the physical and mental worlds, participates in both by manifesting
new essential structures – what Husserl calls “fulfillment”18 – that complete
the modalities of the geometric and physical extensions and of thought.
Let us here briefly clarify some of the meanings more directly related to
our topic that Husserl ascribed to the notion of “fulfillment,” which play a
central role in its phenomenological analysis of the world and of conscious-
ness. Husserl proposes a systematic examination of the distinction between
representations and intuitions, and of the relationship of “fulfillment” which
may hold between them. He gives the following characterization: “psychical
experiences” which do not include their objects in themselves as immanent
contents (thus as present within consciousness) are called “representations”
by him. They merely intend a content or object, which means that it is aimed
at, minded, or referred to with understanding, by means of some contents that
are given in consciousness and are used as representatives of the former. In
contrast, there are certain other “psychical experiences” that do not merely
intend their objects, but really include those objects within themselves as their
immanent contents. They are called “intuitions.” Husserl proceeds to explain
his central concept of fulfillment in terms of relationship between acts of con-
sciousness. We often directly experience an inauthentic or representational
thought of an object finding “satisfaction” in an intuition of that object as
thought of. In Husserl’s words, “The immediate psychical experience of the
fact that the intuited is also the intended shall be designated as conscious-
ness of the fulfilled intention.” The representation finds its fulfillment in the
intuition. Note that a representation finds its fulfillment, not immediately, but
only by passing through other representations that are “closer” to the ultimate
object. The ultimate fulfillment of any representation is the intuition proper
to it. It is purely intuition – a term which expresses the fact that a content
(present to us) bears no representative function whatsoever.
In the fifth Logical Investigation Husserl worked out a theory of intention-
ality that permits consciousness to reach beyond its components and join,
in fulfillment, with a mind-independent object. This theory rests on the dis-
tinction between the “real object” and the “intentional object.” In fact, the
“intentional” object is the same as the “real” object. No act of consciousness
is as a whole about two objects, one in the act and one out. The cardinal ques-
tion of the objectivity of knowledge (or of epistemology) is stated by Husserl
in terms of “the relationship between the subjectivity of knowing and the ob-
jectivity of the content known, or the relationship between act of thought and
fulfilling intuition.” The fulfillment of thought – the “knowing” involved –
is my consciousness of an (any) object. The thought or “empty intention” is
what is fulfilled, and the intuition is what “gives” the thought its fulfillment.
The object of knowledge is the same as the object of the conceptualization
and the object of the intuition, but it is not presented in the same way to the
act of knowledge. One might say that the object of intuition is what would
230
result from a continuous succession of fulfillments, in each of which some-
thing new, that would display and enrich the sense of the object of knowledge,
is added. A fulfillment is one type of “synthesis” of identity or identification.
According to Husserl, it is the identity of the intention or “meaning” in the
conceptualization, on the one hand, and in the perception (intuition), on the
other, that brings the “fullness” of the object through the perceptual act to the
act of conceptualization or mere meaning. This latter is then “filled full” of
the reality of the object itself. It should be added however that the intuition
involved in a fulfillment must be adequate if that fulfillment is to be a case of
certain knowledge, and it is adequate just in case every aspect of the object
as conceived is also directly given. With respect to our perception of an ob-
ject, we have thus to distinguish between those parts of the object which are
present, that is, which pertains to the present perception of the object, and all
those aspects (inner and unseen parts) of this object which are not yet present
to our perception. For, the perception of objects would develop if we were to
proceed, by performing some intentional acts, such as movements, actions,
bodily and mental operations of different kind, to reveal other aspects and
moments of their history and objective becoming.
In other words, there is possible, phenomenologically, a continuous flux
of fulfillments or identifications, in the steady of serialization of the percepts
pertaining to the same object. Each individual percept is a mixture of (pro-
visional) fulfilled and unfilled intentions and intuitions. To the former cor-
responds that part of the object which is given in more or less projection in
this individual percept (indeed any ‘perceptual object’ should be seen as a
topological-differential projection of a physical object with its real and spatial
properties into the multi-dimensional neural space furnished with its physio-
logical substrate and sensory systems), to the latter part of the object that is not
yet given, but that new percepts would bring to actual, fulfilling presence. The
result is a sequence of syntheses of identifications binding self-manifestations
of an object to self-manifestations of the same object.19 As it as been rightly
noticed, “At the lowest levels of fulfillment we have only an abstract descrip-
tive knowledge of objects that does not even reveal their identity. Progression
in fulfillment eventually brings us to a point where the objects are individuated
for us – we know which entities they are – because they are given in terms of
properties that form some significant part of their identity.”20
To put it in another way, progressive fulfillment is correlative of individu-
ative consciousness, and this correlation is made possible by a process of in-
tentional performing accomplishment, or dynamic active perception. Husserl
holds that only perception of objects, not any type of symbolic or even imag-
inative representation, can bring us to this level of comprehension. But it is
only adequate perception that guarantees the existence of the corresponding
object. Adequate perception is given when there occurs a dynamic coincidence
between intention and intuition, which is not merely verification over time but
231
the perceptual determination of sense over time. For example, an identical–
unchanged spatial body, in spite of the fact that continuous perceptual dis-
closing of an object involves variations in sensory input, is a verification of
its identity, as such in a kinetic perceptual series that continuously yields ap-
pearances of its different sides.21 In each case, perceptual content turns out to
have the property of transposability, that is, of being a whole which remains
constant in spite of variations among its parts.
Nevertheless (and here in our opinion is one of the most important lim-
itations detracting from the analyses and conclusions the philosopher has
developed on the subject of spatial perception and the constitution of material
nature), Husserl does not arrive at an understanding of the phenomenal world
as such, other realities aside, as presenting a specific geometry that confers
on it a proper organization and meaning – that is, hence, a semiogeometry
in the sense of Y. R. Thom (see Thom, 1991). Husserl still considers the
physical world and the phenomenal world as being in a way separate. He
does not seem to take into consideration the fact that it is precisely these
appearances that are essential modalities of being through which, in fact, sub-
stances are given form, that is to say a proper existence or an identity. He
wrote:

The geometrical determinations pertain to the physicalistic Object itself;


what is geometrical belongs to physicalistic nature in itself. But this is not
true of the sensuous qualities, which thoroughly belong in the sphere of the
appearances of nature (E. Husserl, Ideas Second book, p. 82).

It is in fact the fundamental motive, we believe, that leads Husserl to an


even more radical philosophical conclusion regarding the opposition between
“geometric concepts” and “descriptive concepts.” The first belong to an on-
tologically formal world, whereas the second pertain to a world of natural
forms and sensible patterns. Husserl thus undervalues the fact that the geo-
metric (ideal) concepts, such as spheres, spirals, polyhedra, manifolds, tensors,
knots . . . , and the morphological (descriptive) concepts, such as shells, trees,
colors, crystals, waves, cells . . . , might be closely interrelated, as indeed they
are. Let us quote him:

Geometric concepts are ‘ideal’ concepts, they express something which one
cannot ‘see’; their ‘origin’, and therefore their content also, is essentially
other than that of the descriptive concepts as concepts which express the
essential nature of things as drawn directly from simple intuition, and not
anything ‘ideal’. Exact concepts have their correlates in essences, which
have the character of ‘Ideas’ in the Kantian sense. [. . .] Descriptive concepts
are morphological concepts of vague types of configuration, which on the
basis of sensory intuition are directly apprehended, and, vague as they are,
conceptually or terminologically fixed. [. . .] The most perfect geometry
and its most perfect practical control cannot help the descriptive student
232
of nature to express precisely (in exact geometrical concepts) that which
in so plain, so understanding, and so entirely suitable a way he expresses
in the words: notched, indented, lens-shaped, umbelliform, and the like –
simple concepts which are essentially and not accidentally inexact, and are
therefore also unmathematical (E. Husserl, Ideas, First book, pp. 201–208).

It is worth noting that at the same time than Husserl was writing the pre-
vious passages, the Scottish naturalist D’Arcy Thompson consecrated an en-
tire book22 to show that the forms of natural patterns and living beings can
be explained by geometric notions and transformations, and that in general
non-organic forms exists save such as are in conformity with physical and
mathematical laws. Let us quote him:

The physicist proclaims that the physical phenomena which meet us by the
way have their forms not less beautiful and scarce less varied than those
which move us to admiration among living things. The waves of sea, the
little ripples on the shore, the sweeping curve of the sandy bay between
the headlands, the outlines of the hills, the shape of the clouds, all these
are so many riddles of form, so many problems of morphology . . . They
have also, doubtless, their immanent teleological significance. (. . .) Nor
is it otherwise with the material forms of living things. Cell and tissue,
shell and bone, leaf and flower, are so many portions of matter, and it is
in obedience to the laws of physics that their particles have been moved,
moulded and conformed. Their problems of form are in the first instance
mathematical problems, their problems of growth are essentially physical
problems, and the morphologist is, ipso facto, a student of physical science
(Introduction).

Husserl, on the contrary, maintains that geometry is not in the least de-
scriptive (i.e. a natural science) and phenomenology (and morphology) are
not essentially geometric. This is precisely where the main weakness of the
phenomenological project lies; because even if geometry is indeed not of a
descriptive nature, the world of phenomena and forms can nonetheless possess
a proper geometric structure as we shall attempt to illustrate in the following
pages.
The foregoing considerations require further qualifications, however. The
development of phenomenological analysis, it seems to us, can be divided
into two fundamental phases. The first phase, which coincides with the work
that culminates in Ding und Raum – Vorlesungen 1907 (Thing and Space,
Lectures of 1907) and which still continues to produce a few echoes in
Phänomenologische Untersuchungen zur Konstitution (Ideas Pertaining to a
Pure Phenomenology and to a Phenomenological Philosophy. Second Book:
Studies in the Phenomenology of Constitution) Husserl seeks to show the dif-
ferent geometric structures that underlie the phenomenal constitution of the
“spatial thing.” From this point of view, we can say that the meaning of the
content of perception is inseparable from the type of geometry presented by
233
the phenomenal field, which, in a sense, conditions the modes of the contents’
appearance. In the Ideen, we witness a change of perspective, whereby the
contents of perception are conceived as separate from the objective conditions
by which they are nonetheless accompanied.
With respect to perception, the correlative distinction with those between
noema/noesis is among the (conscious) acts of perception and their intentional
contents, and the objects represented or “intended” in these acts and the states
which underlie them. Every act has a content, but every content not necessarily
has an object. Moreover, the “object of perception” always transcends the
content of any given act, in that there are always further aspects of the object
which are not in any way represented within it. Likewise, the activity of
consciousness is so rich and complex that it can be parsed into acts, their
parts and moments, in a variety of different ways. According to the view that
Husserl laid out in the Logical Investigations and developed in Ideas I, the
content of an intentional act is its noema. He distinguished between “real” and
“ideal” content of an act. A noema is an abstract entity, like a concept. It is
formed from a sense or content together with a certain “thetic” character such
as that of imagining, perceiving, judging, etc. For the real content of an act
Husserl introduced in Ideas I the term noesis or noetic moment, which is then
correlated with the intentional content or noema of the act. Furthermore, the
“intentional” object and the “real” object are the same thing, and they form a
whole in each act of consciousness. If the real object does not exist, neither
does the intentional one, and if the intentional one exists, so does the real one,
for they are one and the same.23
For all that, it seems to us that Husserl clearly opposes the sensible status
of appearance to the physical status of the perceived object, and even when he
links the two it is only to send us back to an idea first proposed by Helmholtz,
whereby perception is only the sign of real relationships that exist between
physical events. There are, however, profound differences between these two
conceptions.
Helmholtz states that,

Our sensations are indeed effects produced in our organs by external causes;
and how such an effect expresses itself naturally depends quite essentially
upon the kind of apparatus upon which the effect is produced. Inasmuch
as the quality of our sensation gives us a report of what is peculiar to
the external influence by which it is excited, it may count as a symbol
(Zeichen) of it, but not as an image (Abbild). For from an image one
requires some kind of alikeness with the object of which it is an image
. . . But a sign need not have any kind of similarity at all with what it
is the sign of. The relation between the two of them is restricted to the
fact that like objects exerting an influence under like circumstances evoke
like signs, and that therefore unlike signs always correspond to unlike
influences.24
234
For Helmholtz, this trace of resemblance, contrary to contemporary opin-
ion, is not in the least minor, because it is thanks to it that the lawfulness of
phenomena occurring in the real world can be fulfilled. All laws of nature
state that at all times, if the initial conditions within a certain relationship are
the same, consecutive phenomena are produced which, given another rela-
tionship, are identical.

To popular opinion, which accepts in good faith that the images which
our senses give us of things are wholly true, this residue of similarity
acknowledged by us may seem very trivial. In fact it is not trivial. For with
it one can still achieve something of the very greatest importance, namely
forming an image of lawfulness in the processes of the actual world. Every
law of nature asserts that upon preconditions alike in a certain respect, there
always follow consequences which are alike in a certain respect. Since like
things are indicated in our world of sensations (Empfindungswelt) by like
signs, an equally regular sequence will also correspond in the domain of
our sensations to the sequence of like effects by law of nature upon like
causes. [. . .] Thus although our sensations, as regards their quality, are only
signs whose particular character depends wholly upon our own makeup,
they are still not to be dismissed as a mere semblance, but they are precisely
signs of something (Etwas), be it something existing or happening, and –
what is most important – they can form for us an image of the law (Gesetz)
of this thing which is happening (Ibid., p. 122).

Husserl, for his part, writes that, in effect,

the whole essential content of the perceived thing, all that is present in
the body, with all its qualities and all that can ever be perceived, is ‘mere
appearance’, and the ‘true thing’ is that of physical science. When the
latter defines the given thing exclusively through concepts such as atoms,
ions, energies, and so forth, and in every case as space-filling processes
whose sole characteristica are mathematical expressions, its reference is
to something that transcends the whole content of the thing as present to us
in bodily form. It cannot therefore mean even the thing as lying in natural
sensible space; in other words, its physical space cannot be the space of the
world of bodily perception. . . The ‘true Being’ would therefore be entirely
and fundamentally something that is defined otherwise than as that which is
given in perception as corporeal reality, which is given exclusively through
its sensory determinations, among which must also be reckoned the sensori-
spatial. The thing as strictly experienced gives the mere ‘this’, an empty
X which becomes the bearer of mathematical determinations, and of the
corresponding mathematical formulae, and exists not in perceptual space,
but in an ‘objective space’, of which the former is the mere ‘symbol’,
a Euclidean manifold of three dimensions that can be only symbolically
represented (Ideas . . ., first book, op. cit., pp. 128–129).

In short, it must be underlined that for Husserl all things that pertain to the
mental sphere proper do not possess a spatial character, whereas all things that
235
arise from the phenomenal world and that are “filled” by sketchily appearing
sense qualities are possible only spatially. To better illustrate this difference,
he uses the following image: the immanent meaning of a sense quality (color,
for example) relates to the identical (invariable) unity in the same way as the
empirical and physical content of that same quality relates to the continuous
manifold of its various changing sensory aspects. In this way, however, Husserl
cannot think of the “continuous manifold” of phenomenal sketches in terms
of the exact mathematical concept of manifold, which is endowed with a par-
ticular geometric structure whereby it can account for the phenomenological
properties of any sensory field. Instead of considering the concept of manifold
as a mathematical model, a foundation for explaining natural phenomena, he
proceeds to think of it in logical or axiomatic terms; more precisely, in the
sense of the purely formal concept of univocally and finitely “definite” mani-
fold (“definite” Mannigfaltigkeit) or “mathematical manifold in the pregnant
sense of the term.” Indeed,

It [i.e., the concept of ‘definite’ manifold] has the following distinctive


feature, that a finite number of concepts and propositions [. . .] determines
completely and unambiguously on lines of pure logical necessity the total-
ity of all possible formations in the domain, so that in principle, therefore,
nothing further remains open within it. [. . .] A manifold of this type has the
distinctive property of being ‘mathematical, exhaustively definable’. The
‘definition’ lies in the system of axiomatic concepts and axioms, and the
‘mathematically-exhaustive’ herein that the defining assertions in relations
to the manifold imply the greatest conceivable prejudgment–nothing fur-
ther is left undetermined. [. . .] In a mathematically definite manifold the
concepts ‘true’and ‘formal implication of the axioms’are equivalent, and
likewise also the concepts ‘false’ and ‘formally implied as the opposite of
a formal implication of the axioms’ (Ibid., § 72, pp. 204–205).

We tend to interpret this Husserlian thesis in the following manner. The op-
position between “descriptive sciences” (phenomenology, morphology) and
“exact sciences” (geometry, etc.) rests, in our opinion, on an even more fun-
damental opposition: that between real space (in essence, surrounding Eu-
clidean space) whose geometry is founded on determinations of intrinsic
(factual) relations, and space as an abstract manifold (most significantly the
models R 3 and R n ), which is founded on determinations of a purely formal
nature.
Yet, Husserl maintains (in works prior to Logische Untersuchungen) that
the space of physical phenomena (factual space) is a “concretization” (eine
“Konkretion”) of abstract Euclidean space, without seeking to inquire, how-
ever, if this very Euclidean geometry is not also at work in the phenomena
of perception and mental activity. His failure to truly account for such a pos-
sibility is also explained by his stand, against Riemann in particular and in
agreement with Hilbert on this point, that the significance of the theory of
236
manifolds is solely a formal one. Husserl has in mind of course much more
the set-theoretic definition of the concept of manifold given by Georg Cantor,25
as an abstract set of points, than those given by Riemann and latter system-
atized by Weyl,26 according to which a manifold is essentially a geometric
being characterized by abstract relationships, such as coverings, homeomor-
phisms, maps, functions, between the components (i.e., the neighborhoods)
forming the manifold, and endowed with at least one, if no several, metrical
and topological structures to which those relations apply.
In his notes for a lecture (delivered at Göttingen in 1889/1990),27 Husserl
had already openly criticized Riemann’s conception, and especially the idea
that it is possible to reduce the problem of the physical and phenomenologi-
cal constitution of space to the mathematical (analytical) construction of the
general concept of manifold. And yet he excludes the possibility that such a
concept can contain intrinsic and/or real spatial significance. He writes that:

the analytic Riemannian concept of manifold does not offer the possibility
of a veritable intrinsic generalization of the concept of space, but rather, a
simple formalization (einer blossen Formalisierung), which does not allow
for a stricto sensu distinction of any other spatial form (Hua XXI, vol. XXI,
pp. 344–345).

According to Husserl, the true geometric theory remains the Euclidean


geometry, and the value of Riemann’s new mathematical concepts consists
only in their ability for logical generalizations of Euclidean geometry. He
seems to believe in a certain primacy of Euclidean geometry with regard to
new geometric theories. Husserl’s judgment is, as we have shown, mistaken
on both the mathematical and physical sides (see Boi, 1995). It is true that
at the time both Einstein’s theory of general relativity and Weyl’s great work,
Raum-Zeit-Materie, showing the essential role of the concept of manifold in
understanding physical phenomena, had yet to be divulged. Nonetheless, at the
time when Ideen I was being written, the discovery of special relativity (1905)
together with the publication of Minkowski’s Raum und Zeit (1908), wherein
the notion of space–time in the sense of a four-dimensional and metrically
negative (hyperbolic) manifold – so fundamental to physics – is defined, should
have furnished Husserl with the proof that there were other spatial forms,
beyond Euclidean space, that lent themselves to concrete interpretation.
By the concept of “manifold” (Mannigfaltigkeit), Husserl designates a
formal domain of objects as the objective correlative of a formal system
of axioms. He does not allow that every manifold, as an abstract mathematical
concept, can be the “support” for a particular space in which a certain kind
of geometry can be carried out (which does not mean, naturally, that a geom-
etry may not exist independently of substrate space – in fact, all axiomatic
geometry only “exists” in this sense). For Husserl, spaces that are obtained
through a generalization of Euclidean space, in the framework of manifolds
237
theory, are only so many category-specific forms. In this respect, it is interest-
ing to recall Husserl’s frequent insistence on an important distinction between
a “Euclidean manifold” (or any other abstract manifold, for that matter) and
“space”. In an 1887 text,28 he underlines this distinction in more or less the
following terms: the concept of time differentiates itself from that of possi-
ble “orthoid manifolds” (“orthoide Mannigfaltigkeit”) by virtue of substance
(Stoffliches) and “concretization” (Konkretion). Whereas there exists an end-
less number of possible orthoid manifolds, there is only one time; similarly,
there exists an endless number of possible Euclidean manifolds, but only
one space. Euclidean manifolds are only one among many types of possible
manifolds, just as the number two is just one among many types of possible
numbers in the series of natural numbers. They are purely formal types, as
they are wholly pure regardless of the substance. Conversely, the type that
constitutes colors, formed out of the smallest nuances (differences) among
them, even when allowing for a distinction (i.e., between primary colors and
the varieties of secondary colors) cannot be determined in a purely conceptual
way. Given that the spatial point, as such, is the concretization of the several
points of a manifold and that geometry has nothing to do with spatial physical
determinations, it is then understood that geometry can be differentiated in
terms of its systematic (axiomatic) form of the theory of Euclidean manifolds
only if one were to replace elements in general by elements that are properly
spatial. If we then make abstraction of the matter of space and consider it as
a pure form, we will then have ipso facto the theory of Euclidean manifold,
in which any given points are objects that are determined by certain formal
properties characterizing them.
In another text,29 Husserl relies on the analogy between the category “qual-
ity” color and the formal concept of “manifold” to show that just as one pro-
ceeds from a manifold, through simple formal determinations, not to space but
only to a Euclidean manifold, one cannot proceed from quality to color or to
any other given color. Purely formal operations are insufficient in both cases:
“fulfillment” of a different kind must intervene. In other words, that which
transforms an Euclidean manifold into space relies as little on formal princi-
ples as the transformation of quality into color. Indeed, as Husserl points out
here and later expands upon in Ding und Raum, the spatial figure is a moment
in the intuition inasmuch as the latter determines an intrinsic constitution of
the phenomenal object as well as its color or shine. In which case we could
think that,

as the category ‘quality’ corresponds to color (that is to say to the kind of


intrinsic constitution that is reserved in an identical and undivided manner
to the objects in its entirety, to an entire part of the object), similarly, the
category ‘manifold’ corresponds to the ‘spatial’ (Raümlichen), and does so
in the particular formation (Ausgestaltung) we call homogeneous manifold
and, more specifically, three-dimensional Euclidean manifold (Ibid).
238
From a certain point of view, this position may seem completely justified
once we admit not only that the notions of “space” and “sense quality” are of a
supreme logical genus but also that they are ontologically indivisible “primary
substances.” We will present the following arguments to contest this point of
view:
(a) It is the real that is geometric and not geometry that is real; so that even
phenomenal space can present interesting geometric properties, which
indeed seems to be the case.
(b) Nothing bars us from considering that real space underlies numerous
abstract geometric “objects” which we shall call “manifolds” (endowed
with sound mathematical properties), and which can serve as models for
substratum spaces in which a set of sense qualities is defined and its
evolution pursued in time. In other words, the concept of manifold, and
especially of Riemannian manifold, despite its ideal genesis, is a principle
of intelligibility suited for the explanation of certain fundamental physical
phenomena as well as their modes of manifestation.
Husserl’s conclusion which leads, as we have just witnessed, to a dichoto-
mous conception of space and of manifold, is all the more surprising be-
cause elsewhere, especially in Ding und Raum, he repeatedly displaces the
concept of manifold outside its original mathematical context in order to
apply it to the phenomenal domain of the constitution of the res. At any
rate, it is thanks to the concept of manifold that he can characterize the
structure of the phenomenal sensory field (particularly the visual field) in
which the unilateral appearance of a “thing” is constituted through identi-
fication, and the form of the “spatial body” is fully defined. In short, he in
some way projects those characteristics pertaining to the concept of manifold
onto the general properties of sensory fields (visual, tactile, and oculomotor)
which he regards as nothing else but mutually correlated formal systems of
reference.30
Furthermore, it remains to be seen whether these fields are independent with
regard to the structure and the organizing laws of three-dimensional objective
space. The same applies to the finite and closed system of axioms and rules
that define the domain of objects of a manifold and their correspondence to a
unified system of field that is permanently organized and controlled through
an a priori lawfulness which implies the unity of kinesthetic/intentional moti-
vation. In such a way, the two systems achieve continuous global synthesis. It
is imperative to grasp this nodal point. Let us suppose a series of images (of the
visual field) i p continuously unfolding in time in a temporal coincidence with
a continuous series of kinesthetic movements (and sensations) kq . This conti-
nuity of images (a series of “fulfillments”) is a linear manifold, extracted out
of a multi-dimensional manifold of possible images – in an analogous man-
ner, so to speak, abstract Euclidean space R 3 is obtained from n-dimensional
239
Euclidean space R n (incidentally, any surface or abstract manifold is defined
by an atlas of maps where each map is an open set of R n ),

and, like the latter, it still includes infinitely many other linear manifolds of
images, and each one, according to its determinate type, is encompassed by
the determinate general type of the total manifold. This one is of the same
strength (Mächtigkeit) as the continuous manifold of possible k’s. Every
actually elapsing double manifold of images and of k’s is united through
the unity of the continuity of apprehension, which functionally unites the k
and i belonging to every temporal phase into an apprehensional unity (into
an appearance) and unifies the appearances into a temporally flowing total
appearance (TS, pp. 157–158).

It is precisely this kind of process which underlies the constitution of the


spatial thing. Indeed,

The unity-consciousness which develops in this continuity of appearance,


along with the essentially concomitant continuity of the k-motivation, posits
the unity of the thing; i.e., it constitutes the thing. It belongs to the sense of
this unity, which we call the thing, to be unity in a manifold of appearances,
in a continuity of appearances of a determinate, ideal, infinite type. And
the thing is given in every actually streaming piece of this continuity. The
thing is also given in every appearance of such a piece, and in each one
the consciousness of the identity of the fulfillment is alive, even if it is a
changed appearance . . . But it pertains to the essence of this givenness to
leave open infinitely many possibilities of new givenness in a determinate
way as motivated possibilities (Ibid., p. 159).

When attempting to formalize the foregoing, we can affirm that these sets form
what could be called a stratification, in which the layers are series of images i p
or the kinesthetic series kq defined in the very large space P (which we shall
suppose is also finite for the sake of simplicity) which is formed of a finite
number of domains Ur belonging to the objective space of perception and
isomorphic to open sets of E n , such that P = ∪ Ur ; between these two types
of series, a one-to-one mapping can be established, which is to say, that one can
differentially express one system as a function of the other: i ap = f (kqb ), where
a, b = 1, . . . , n, and vice versa. The stratified sets form a continuous series
S1 → S2 → · · · → Sn and yield a space global product  :  S P ∪ U(i,k) .
Such formalism seems to hold a certain importance inasmuch as it is suited
to describe the “primitive beings” from which space is constituted. We come
across an idea here which René Thom set forth some years ago, albeit in
a somewhat different context. In his case, it was a matter of showing that
the presence of a metric in the sensorimotor spaces of animals is an obvious
conceptual necessity. Nevertheless, the problem is basically the same: how to
explain the constitution of perceptive space, that is of a system that ensures
the coordination of different sensorimotor spaces. If we disregard as unlikely
240
the idea that sensory space, in the case of our species, is genetically entirely
constituted at birth, then we must allow for the role played by a certain number
of “primitive spaces”: the postural space of the organism’s entire gamut of
posture; the organogenetic spaces, which, in some cases, lead to the formation
of organs that are very controlled metrically, such as the eye or the muscles;
and other physiological spaces. Thom’s ideas on this important issue point in
the same direction as the analyses of Husserl. Indeed, according to Thom,

naturally, all these spaces have an Euclidean (multi-dimensional) spatial


structure and many have a metric structure. In fact, we have before us an
operation of ‘synthesis’ which requires the elimination of a great number of
superfluous parameters: space is not ‘constructed’ as an articulated piece,
by the composition of pieces, but rather, as a quotient by identification of
the product of a large number of physiological spaces, most of which are
endowed with all desirable structures.31

5. The Distinction Between “Orthoid” and “Cyclical” Manifolds, and


Its Role in the Constitution of the Objective Space of Perception

Let us return to Husserl’s distinction between “orthoid manifold” and “cyclical


manifold.” For Husserl such a distinction, as he shows, can lend itself to
a concrete interpretation because its involvement in space coincides with
the conversion of the oculomotor field into an Euclidean manifold. First off,
the oculomotor image (of an object) is constituted by its separation from
other objects by virtue of a qualitative discontinuity. The object is a closed
configuration (or shape) that qualitatively detaches itself from its background
in the visual field. But the image, as such, results from different moments that
establish themselves in such a way as to form a continuous merging unity. In
fact, these moments amount to expansion, with its reliance on proximity and
distancing as properties of modification and on self-rotation to dissimulate and
reveal the object. Pure distancing, therefore, is a linear modification; which
is to say that the motivating circumstances vary in a fashion which is linearly
“orthoid” to the infinite. From a phenomenological point of view, this means
that the object can eventually leave the visual field and go beyond the horizon
(the point at infinity). Pure turning is a cyclical modification, in the sense
that

the kinesthetic circumstances vary cyclically, and in the system of pure


modifications of turning they bring back the turning series of images (TS,
p. 212).

By way of the cyclical character of appearances (the object’s faces), the


closed external surface (the boundary) of the body is constituted. In this
way, the set of continuous rotations generates the surface of the body and
241
constitutes a two-dimensional cyclical manifold of points. With respect to
expansion, Husserl remarks that

mere expansion is not a modification that is related to the change of mere


orientation in the oculomotor field as are two powers which can be combined
according to the principle of the parallelogram of powers (Ibid., p. 213).

In other words, the expansion of an object in the oculomotor field does not
depend on its orientation in space. An object can be said to be oriented if
any figure undergoing a displacement or a rotation keeps invariant in the
oculomotor field.

For instance, in the case of ‘displacement’, every point in the field maintains
the same linear transformation of coordinates: every determinate distance
undergoes the same mere displacement in parallel. The reciprocal orien-
tation of the points toward one another is the same in the one location of
the figure and in the others. As regards expansion, on the other hand, the
points do not retain their reciprocal orientation (Ibid).

It remains, nonetheless, unfortunate that Husserl, given this reasoning, was


unable to arrive at more general conclusions. But at one moment in particular,
he does account for a larger notion of the spatial, albeit within the boundaries
of the privileged position that he constantly assign to Euclidean space. It
is in Ding und Raum (Thing and Space), that he discusses the systematic
constitution of space. In the first place, bodies are said to be “embedded” in
space, which for Husserl means:

(i) Considered as a sensible schema, there is a spatial corporeality (or con-


figuration) over which sense qualities stretch. This announces another
important theme, namely the relationship between spatial extension and
sense qualities; these are necessarily merged with a certain extension,
which is constitutive of the substrate (see infra). Every such sensuous
schema counts with one or more variable places, “such that; given the
identity of the schema, a closed system of locations (orientations) is
possible idealiter” (TS, p. 297).
(ii) The sense of touch and the sense of sight constitute a body as a sen-
sorial schema, and every sense is a sense through an apperceptive (ap-
perzeptive) conjunction of the corresponding sense data with kinesthetic
data.

With regard to the relationship between spatial extension and sense quality,
it is important to note that the quality participates in a certain manner in the
modification of the extension. In fact, the quality is affected by modifications
of the extension, notwithstanding the fact that the kind of modification which
242
is proper to it remains independent. Let us think of the color that covers a
certain type of physical surface, or better yet, the external of crystalline layer
– the iris – which dilates or contracts in accordance with the (qualitative)
nature of the luminous source it receives, whereas the zone at the center of the
retina – the fovea – on which the focal point is imaged remains immobile.32 It
therefore follows that the two contents, by their very nature, cannot be disas-
sociated, that they form in one way or another a unique content of which they
are only partial contents. But it is not a merely quantitative matter here, in the
sense that each time the “quantity” of the extension is either increased or de-
creased the same will apply to its quality. Rather, things have to be considered
qualitatively. From this vantage point, it would be entirely unintelligible for
the quality to diminish in such a way and erase itself by a simple diminution
and disappearance of the quantity without modifying itself in its own manner
as a quality. At least, insofar as perception is concerned quality and quantity
cannot be autonomous contents; that is to say, they can neither be separated nor
exist in mutually independent representations. We must admit, consequently,
that, in this sense, space (spatial extension) is as primary as quality and that
its perception is equally direct and generally occurs simultaneously. In other
words, any space is immediately given together with and in the representa-
tion of quality. Insofar as it is phenomenal space, its characteristic property
consists in the fact that its continuity corresponds to the continuity of qual-
ities, and that the diverse qualities (visual and tactile), devoid of connection
in themselves, are thereby given unity. Let us suppose, with Husserl, that the
global coloration of a surface is composed of color elements, of qualities that
partly propagate themselves from one point to another, and that present, in
part, discontinuous leaps in isolated points,
Then these qualities are ordered in the expanse [Ausdehnung], and in it they
fuse into unitary colorations which have their own unitary form of order.
Coloration is the way the points of the expanse are colored in their order
and the way they fuse, in virtue of this order, into an unity and present a
unity as a qualitatively conjoined one. But the conjunction does not reside
in the colors in themselves; as ordered differently within the same expanse,
they result in different units of coloration (TS, p. 304).
But the liaison does not exist in the colors themselves, but rather in the kind
of spatial relation that unites the extension to the color that covers it. To
fully understand the foregoing, we must underline one fact of general import,
namely that spatial perception confirms the hypothesis according to which the
visual field does not extend itself without regard for the organized character
(into regions) of the visual field itself, but rather, that the homogeneous field
has a tendency to increasingly manifest itself as being uniformly unitary. Such
a unity is moreover composed neither of points nor of the relationships between
them. Rather, it is a matter of pregnant forms: allowing us to see limited regions
spontaneously detaching themselves in a sharp and increasingly accentuated
243
manner from the rest of the field, while the whole field continues to preserve
the aspect of a phenomenal “nexus.”33 It is, in fact, possible to propose an
even more general thesis (which we shall not justify presently), namely that
psychophysical phenomena that produce themselves in the optic system share
(almost) the same general properties with physical spatial configurations (see
on this subject Boi’s book, 2004). We shall proceed now to the constitution of
space proper. It includes the following levels, which have been well described
by Husserl:
1. Visual space (of the first level) is oculomotor space. Here only the eye
moves. This space is restricted in a fixed manner. No sooner do bodies
appear in it than they disappear. This space contains a privileged null-
point to which the basic position of the eyes – the frontal position – is
kinesthetically coordinated. The two principal axes of the coordinated of
the oculomotor system – the above–below axis and the right–left axis –
intersect in this frontal position.
2. A closed, two-dimensional space is constituted by the rotation of the head
around its basic axis, while the remainder of the body is in its normal
position, a position that is supposed to remain motionless and fixed. The
oculomotor visual space is an enduring form, which, however, receives a
new index with every position of the head. We then have a new kinesthetic
system as a new line of motivating data. This new space has no limits on
the “right and left;” it is a closed space. What functions here as the basic
system of orientation is not an “intersection of axes” but a closed line
of coordinates in the right–left direction, as an abscissa, and an unclosed
line in the above–below direction.
Already at this level, we are no longer dealing with Euclidean perceptual
space, properly speaking, but rather with a two-dimensional space (a
surface), which globally differs from it but is locally equivalent: strictly
speaking, this space has the form of a cylinder. As Husserl very rightly
observes:
The spatial field would be delimited above and below, and we would
have a null-line, the closed abscissa axis and two parallel lines
y = +b, y = −b. For a = 0, we would have a length with two
opposite directions (traversal from below to above and from above to
below) (TS, p. 267).
3. If we consider the whole cephalomotor space, but imagine that everything
that constitutes depth has become imperceptible, that is to say lacking in
all proper kinesthetic movement (rotation in opposite directions), we will
then have, with the supposition that the mobility of the head is idealized
in an appropriate manner, a closed spherical space.
4. So far, all bodies are two-dimensional (“like-surfaces beings”), or in the
best of cases, “spherical beings.” This would constitute a homogeneous
244
two-dimensional Riemannian space. A third dimension, depth, and a
three-dimensional corporeality would have no sense. But those are con-
stituted when groups of movements, or groups of kinesthetic data, are
coordinated to image changes of a new kind. That is the way that visual
space as Euclidean space or as complete visual space should form.

In this way, Husserl arrives at a twofold conclusion. On the one hand, he


affirms that the space of vision is a Euclidean spatial form, albeit with the im-
portant consideration that Euclidean space is only constituted for surrounding
bodies and not for the body proper or the “null-body” (that is to say, a body
that is limited in each of its movements). On the other hand, he holds that
space is, nevertheless, not entirely Euclidean.

Geometrical space in the Euclidean sense is a formal schema for all possible
corporeality, and there is no difference here between the null-body and
the environing bodies. All bodies are equal, the presentation in the null-
appearance is equivalent to any other presentation, and movement and rest
in the geometrical sense are of the same type for every body presenting
itself in any ‘accidental’ orientation. On the other hand, the designated
visual space does have those distinctions, and they can be dissolved only
if the movement of the corporeal body in space is constituted, specifically
as equivalent to any movement of other bodies (Ibid., pp. 337–338).

In the following pages, we shall try to elucidate to what extent such a


conclusion is justified. Let us limit ourselves for now to noting that such
a conclusion rests on the distinction, outlined above, between phenomenal
(perceptual) space and abstract geometric space.

6. Mathematical Properties of the Visual Field, Continuity of Intuitive


Space and Geometry of Sense Qualities (Colors)

We would now like to tackle three questions that, in the present context and
more generally in any context relating to spatial perception, impress us as
being fundamentally significant. We will restrict our present consideration
to the ideas of Husserl and Weyl. (i) What is the geometry of visual space
and furthermore, what are its differences with regard to physical space on
one hand and geometric space on the other? (ii) How can we characterize the
property of continuity in the case of sensible or intuitive space? (iii) What is
the relationship that exists between spatial extension and sense qualities? This
is the perennial problem of the relationship between “primary and secondary
qualities.” These questions are, in fact, closely linked in several ways.

1. Let us begin, therefore, by examining the characteristics of the visual field.


As Husserl, among others, has shown, we can speak of a manifold that
245
is two-dimensional (the term is used here in its strictly geometric sense),
congruent with itself, continuous, simply homogeneous, finite, and of
course delimited; it has a limit (a border), which has no beyond. We can say
that topologically the visual field is an open set that is bounded. The visual
field is not part of objective space – as a surface belongs to Euclidean
space R 3 –, just as points and lines in a field are not points and lines in
objective space and do not share any kind of spatial relation with spatial
points and lines. A certain homology must be preserved, nonetheless,
between the objects that exist in ordinary space and the corresponding
images that form themselves on the retina in the visual field, despite the
particular kind of deformation they experience. But let us, for the time
being, leave this rather important aspect aside. If the visual field is a two-
dimensional manifold, then this means that each part or fragment of the
field is circumscribed by limits (borders) that depend on it and that are, in
turn, continuous manifolds themselves, hence themselves subject to frag-
mentation in such a way that their fragments will share common margins.34

Let us try a purely mathematical formalization of this fact. For the pur-
pose, let us suppose the visual field be a special case of an (n – 1)-sphere
imbedded in S n , we can then apply the following Jordan–Brouwer sepa-
ration theorem.

THEOREM. An (n – 1) sphere imbedded in S n separates S n into two


components of which it is their common boundary.

Proof. If B ⊂ S n is homeomorphic to S n−1 , then H0 (S n – B) ≈ Z, by


corollary bellow. Therefore, S n – B consists of two path components. Since S n
– B is an open subset of S n , it is locally path connected and its path components
U and V , say, are its components. Clearly, B contains the boundary of U and
of V . To prove B ⊂ U ∩ V , let x ∈ B and let N be a neighborhood of x in
S n . Let A ⊂ B ∩ N be a subset such that B – A, is homeomorphic to I n−1 .
Then H (S n – (B – A)) = 0 by the lemma bellow, so S n – (B – A) is path
connected. If p ∈ U and q ∈ V , there is a path γ in S n – (B – A) from p to
q. Because p and q are in different path components of S n – B, γ meets A.
Therefore, A contains a point of U and a point of V . Hence, N meets U and
V , and x ∈ U ∩ V .

COROLLARY. Let B be a subset of S n which is homeomorphic to S k for


0 ≤ k ≤ n − 1. Then

0 q = n − k − 1
Hq (S − B) ≈
n

Z q =n−k−1
246
LEMMA. If A ⊂ S n is homeomorphic to I k for 0 ≤ k ≤ n, then H (S n −
A) = 0.

A related result is the following Brouwer theorem on the invariance of


domains.

THEOREM. If U and V are homeomorphic subsets of Sn and U is open


in Sn , then V is open in Sn .

Proof. Let h : U → V be a homeomorphism and let h(x) = y. Let A be


a neighborhood of x in U that is homeomorphic to I n and with boundary B
homeomorphic to S n−1 . Let A
= h(A) ⊂ V and let B
= h(B). By previous
lemma, S n – A
is connected, and by previous theorem, S n – B
has two
components. Because
S n − B
= (S n − A
) ∪ (A
− B
)
and S n − A
and A
− B
are connected, they are the components of S n − B
.
Therefore, A
− B
is an open subset of S n . Since any y ∈ A
− B
⊂ V
, V
is open in S n .

No one will question that what was suggested by Husserl is the same
method ordinarily used to define the spatial continuum. It states that a por-
tion of space, the surface that limits that portion, a portion of that surface as
well as the line that limits that portion are structures that are endowed with
the property wherein the totality of the points situated inside is arithmeti-
cally constructed as a three-dimensional set of real numbers. But now, the
limits cannot be fragmented (at least, in a discrete manner), they are simple
elements of the extension and of the “points.” The limits of the fragments
(domains) of visual fields are manifolds of continuous points, whose parts
are bounded by points; that is to say, they are lines. In this way, the cohesion
of the field is nowhere broken. Each part of the field can be fragmented, into
an interior and an exterior, and we can move by continuous paths from the
interior of each domain into the interior of any other different domain. In
other words, these parts are connected domains of the global visual space
forming a mathematical structure-like connection, and the paths that allow
the joining of each point of one of these domains to each point of any other
domain are reciprocally isotopic. The final term, the one that can neither be
fragmented nor delimited, is the point. With respect to the latter, Husserl asked
himself

whether in principle a fragmentability (Zerstückbarkeit) in infinitum does


not repose in the essence of the field, or whether the de facto fragmentation,
which leads to minima visibilia, provides in them the essentially ultimate
247
elements, thus whether points and visual atoms are one and the same (TS,
1997, pp. 140–141).

He continues,
To answer this question we would needs to attend to the essential similarity
within the visual field itself, in the great and the small, in the intact and
the partitioned, in displacement and rotation. It is obviously this imma-
nent similarity which, as an evident generic sameness, founds the trans-
ference of the essential relations that are located in, as it were, the macro-
scopic realm to the microscopic ‘atoms’ residing beyond all partitioning
(Ibid).
We would tend towards a different answer to this question. Nothing authorizes
us to assess the “points” of the visual field and the “minima visibilia” as one
and the same thing; for what is perceived at a given moment on the retina as
a point is neither necessarily a real point nor a (mathematical) ideal point,
but the limit beyond which our visual perception cannot reach – either in the
direction of the infinitely small or of the infinitely large. Hence, even in our
macroscopic world, what we may see at a given moment as a point, regardless
of whether or not it is a point that is projected on our retina, could completely
change in extension and appearance if we were to approach it and look at it
more closely, or if we were to use other means that focus and sharpen sight, etc.
As the most intuitive daily experiences confirm this fact, we shall not linger
on it unnecessarily. What is of greater importance, however, is to highlight a
certain number of other points that Husserl particularly insisted on.

(i) A content or an object from the field can change in a quasi-material


fashion depending on quality, degree of brightness and saturation even
if its size relations, extension and position remain unchanged.
(ii) At the same time, however, a content or an object from the field can either
change only with relation to its metric relations, or else quantitatively and
qualitatively at the same time. Quasi-displacement and quasi-rotation as
well as quasi-contraction or expansion and quasi-distortion, and so forth
are thereby made possible. In this way, when the sharpness of qualitative
discontinuities is attenuated, the boundary of an object can become vague.
(iii) When objects move over the boundary of the visual field, significant
qualitative changes of the visual forms occur. This is a proof of the
important role played by what in topological terminology are known
as “boundary conditions” in perception. In this same context, let us
recall that changes that are objectively designated are accommodation
changes; it is especially thanks to this innate physiological mechanism
that we can have a more or less accurate perception of patterns and
objects.
248
With respect to this subject, there is an important point to clarify. The above
changes must not be interpreted in an empirical sense; that is, given qualitative
changes, an image shifting towards the edge of the visual field is not an object
that is really in movement. We see by this that we are not simply dealing with
an “objective” representation here of a direct, term for term account of objects
from the exterior world by our visual system. A type of perception intervenes
at this level, which is representational and already partly symbolic in nature.
The objects of the exterior world are almost not perceived as such, but are
seen in accordance with the type of structure that characterizes the different
peripheral and central perceptual systems and the physical and physiological
laws they obey.
2. Let us confine ourselves to the solid and in some sense concrete ground
of the spatial; it becomes necessary, despite the flux of qualitative and
quantitative changes, for the “thing” to not cease being identical to itself:
therein lies its essence. To that end, it is then necessary for it to preserve its
essential property: continuity. All continuity that unfolds in a real temporal
flux “implies” an identity, i.e.,
there exists the ideal possibility of carrying out the continuity in the
consciousness of unity, thus of intuiting the identity of an object in the
stream of its phases and of determining it as given. Thereby we also have
the possibility of unifying, in a synthetic consciousness of identity, the
phases extracted analytically from the unity as well as the possibility of
raising them to the evident consciousness of the identity of the ‘object’
‘presenting itself’ in these phases. That is an essential law, valid for all
continua elapsing in the pre-empirical stream of time. [. . .] The sound,
which changes, is the unity in the streaming and changing manifold. The
sound is an adequate givenness in the consciousness of unity but is not
a really (reell) immanent givenness; it is a transcendence [. . .] given on
the pure ground of immanence. [. . .] The sound changes its pitch, e.g.,
or its timbre, or the one remains unchanged while the other changes.
Obviously the moments of the individual phases participate in the sub-
stantialization; they receive unity through the unity-consciousness that
binds them together: the pitch of the tone is what is identical in all the
pitch-moments present in the phases of the sound (TS, pp. 141–142).

Nevertheless, as Husserl emphasizes,

Yet no matter how important a piece of the Objectivation of the


thing (Dingobjektivierung) is indicated by the words ‘identity in the
continuity’, still that is not the identity of the thing, and it remains an
enormous task to penetrate through to that identity (Ibid).

The problem, here, is of capital importance to the comprehension of what will


follow – and especially, for appreciation of Husserl’s conception. It is a ques-
tion of distinguishing, at one and the same time, between phenomenological
spatial continuity, of the “first degree,” and ideally posited and scientifically
249
determined mathematical (hence, of a “superior degree”) spatial continuity;
and also of the gap between the identity of the real as a thing (sachliche Iden-
tität) and its objective identity, or, put in other terms, between descriptive and
scientific knowledge. Husserl’s true aim is not to establish a hierarchy between
these two levels; his greater concern, rather, is to arrive at an understanding
of the transition from one level to the other, and of the close link that makes
the two levels, in a sense, so essentially inseparable. This gap together with
the impossibility of confusing the identity as object with the objective identity
depends upon the understanding of the identity of sound, to take up an earlier
example, as an identity only in the actual phenomenological flux which cannot
be carried beyond it. However, it is important to see together with Husserl, that

The thing, however, is in and with the stream not only of its actual changes
but also of its possible changes, and the latter are indeed infinite, though
firmly delimited. And the empirical thing is what appears as identical in
the stream of pre-empirical occurrences, but it is not something identical
‘residing’ in them, something that could be adequately drawn out from
them (Ibid).

But are we to conclude from such reasoning that identity is of two kinds:
the identity of the empirical thing that remains unchanged in the flux of
its own changes, and the identity which “resides” within these changes and
that can eventually be adequately determined? As regards the empirical or
phenomenal thing, Husserl excludes the possibility that two kinds of things
can be perceived: the external things and the immanent things. Two thoughts
come to mind here: one concerns the relationship between perception and
form, and the other to the difference between a phenomenal continuum and a
mathematical continuum.

A. First, one must not exclude that the type of identity of the thing which
seems to us to be always identical to itself is, in fact, very different in
nature from the type of identity that we would grant it if we were to consider
it as a physical object, a chemical body, a biological “being,” or even as
a mathematical entity or a formal category, etc. Each of these modes of
identification obeys a precise criterion. For instance, two “objects” (or
two theories) that can be easily defined as identical or equivalent from
the mathematical point thanks to the concept of isomorphism35 can no
longer be equivalent from the physical point of view. In this way, there
exist different levels of structure, for the most part invisible, that we can
associate to the same object; each of these levels generally is sufficient to
characterize and define it as identical. For example, we can identify a crystal
by its configuration, as well as by its (internal) structure or by its form,
which includes the two other levels but which, incidentally, also amounts to
more than the two together. We shall forego mentioning here the important
250
consequences that such a distinction carries. For the moment, we shall
only underline the impossibility of perceiving the different phase changes
that the object crystal may have undergone, and which all correspond to a
particular internal constitution of the object which totally differs from all
others. Does it make sense, therefore, to assume that the object remains
the same object? It would seem not, given that the form, the structures
and properties by which the object (or body) is characterized are destroyed
as a consequence of the changing of phase. With regard to this process,
let us recall that one of the most important phenomena at work here is
the one known as the symmetry breaking. For example, the solid state of
crystal is characterized by a breaking of the symmetry of the Euclidean
group of translations, and we are left only with the subgroup which leaves
invariant the fundamental cell of crystal.36 This breaking is not, however,
always irreversible. Actually, more often than not, it is possible, under
certain physical conditions, to return to the initial state of a given material
object, in such a way that the lost identity of the object is restored, so
to speak. Such an operation relies on the possibility of reconstituting the
fundamental group of characteristic symmetries of the physical object.
B. Though we may have digressed somewhat from our subject by way of these
last observations, we can re-establish a link to our discussion of Husserl
by turning now to the notion of continuity. In the following, we shall rely
primarily on ideas developed by Weyl, in a number of his works.37 Let us
recall that Husserl had made the distinction, with regard to the characteriza-
tion of the identity of the empirical thing, between its “actual changes” and
its “possible changes.” Although he does not pronounce himself explicitly
on this matter, he, nonetheless, implies that these two types of changes
underlie two profoundly different conceptions of continuity. According to
one, continuity is phenomenal in nature and serves to determine the total-
ity of the actual changes of the appearance of an object; these moments
are held together in the consciousness of immediately given (immanent)
unity and occur in simultaneity with the act of perception. According to
the other, continuity, wherein the infinite is permitted to intervene into the
kind of possible (virtual) perceptions, cannot be enclosed within the limits
of an immediately identical (unchanged) unity. This identical unity of the
object could eventually be given, but in an altogether different manner and
namely by an a priori mathematically defined process. This difference,
which Husserl points out on more than one occasion, not unambiguously
to be sure, must be understood in relation to our earlier discussion of the
concept of space as a continuum of real phenomena – “filled” in part by
the ostensible and physical properties of body surfaces present herein, and
partly by the sense qualities of these same body surfaces – and space as
a continuum defined in terms of formal category: the “form” of the Eu-
clidean manifold or of any n-dimensional manifold, be it Euclidean or not.
251
One would, in fact, have to ask if it is not precisely this type of continuity,
inseparable from the mathematical concept of infinity, that is at work in
the kind of scientific explanation which we already discussed with regard
to the phase’s changes.

Let us now turn to Weyl’s thesis, which, in effect, is at least a twofold one. In
the first place, he affirms that it is impossible to base the intuitive (phenome-
nal) continuum on the mathematical continuum, and that, in turn, mathematics
is not based on the intuitive continuum, but rather on the category of natural
numbers. In the second place, he maintains that the two sorts of continua
are essentially different in nature, in the sense that one cannot, for example,
base a pure theory of time on the concept of numbers. In fact, this leads to a
notion that an autonomous theory of the continuum is conceivable, indepen-
dently from number theory (arithmetic). There is little doubt that Husserl’s
influence is at work here. Let us summarize the most important points of his
reasoning.

(i) The continuity of a function is a transfinite property, that is to say that the
fact of knowing if a function is continuous or not depends on the principles
of definition that we apply not only from the basis of natural numbers, but
also from the other sets of numbers that we can obtain from these and with
the help of these principles. These form, in contrast with finite theories,
an “open” system, and the continuity of the function also remains open.
The continuous function can, however, lose this property if the principles
of definition were to undergo an extension, and consequently, other new
real numbers were to be added to the already present ones.
(ii) Consequently, it is impossible to compare the continuity we ascribe to
a mathematical function with intuitive continuity. Hence, to say that a
function represents the position of a material point as a function of time,
as in x = f (t), is equivalent to the conceptual assertion that the function
is continuous, or that, for all real values of its argument belonging to a
certain interval, the function itself only takes on values belonging to a
certain region. Let us now attempt to translate this idealized proposition
into the following intuitive finding: “I see this pencil lying before me on
the table throughout a certain period of time.” According to Weyl, this
is a perception whose content is so to speak immanent, and it would be
patently absurd to think that it could be undermined by an extension of
our principles of definition,

as if new moments of time, overlooked by my intuition, could be


added to this interval, moments in which the pencil was, perhaps, in
the vicinity of Sirius or who knows where. If the temporal continuum
can be represented by a variable which ‘ranges over’ the real numbers,
252
then it appears to be determined thereby how narrowly or widely we
must understand the concept ‘real number’ and the decision about
this must not be entrusted to logical deliberations over principles of
definition and the like (Weyl 1918, Dover edition, op. cit., p. 88).

It is not at all the same in the case of intuitive perception. Here, continuity
is essentially the characteristic attribute of perception. An object is per-
ceived in the “continuity of identity,” and it is in its constitutive continuity
that the object phenomenon reveals itself and not the opposite. Whereas
here, the definition of temporal continuity, which would have no proper
existence as such, is made possible by the concept of real numbers.
Brentano had already made some very profound remarks on this sub-
ject. In a work that we understand as seminal,38 he insists on the fact that
the concept of continuity cannot be obtained through a combination of
attributes that arise out of diverse intuitions and experiences, but rather
by making abstraction of homogeneous intuitions (einheitliche Anschau-
ungen). He states that continuity is a primary given of intuition. By this
he means that each one of our intuitions, those pertaining to the exterior
world as much as those that belongs to an accompanying interior percep-
tion and by extension those of memory as well, manifest themselves to
us in the form of a continuity. The continuous makes an appearance in
each one of our intuitions, in such a way that its concept cannot be cre-
ated by a complicated (and discrete) combinatory process, but directly
by intuitions thanks to a simple abstraction. Brentano thinks that col-
ors, sight and our very selves, in so far as we see phenomenal qualities
succeed one another continuously in time, are just so many real examples
of continuum.
(iii) To further elucidate this last point, Weyl refers to the difference between
phenomenal time which is an intuitively given continuity, and the concept
of number. To establish a relationship between intuitive time and math-
ematical concepts (numbers), it is necessary to fix a strictly punctual
“now” and to discern time-points (Zeitpunkte) in phenomenal duration.
Each pair of time-points would obey the, say, arithmetical earlier/later
binary relation. Two time-points A, B of which A is the earlier define
a time span AB; into it falls every time-point which is later than A, but
earlier than B. The experiential content which fills the time span AB
could “in itself,” without in any way being other than it is, fall within
some other time; the time span which it would fill there is equal to the
span AB. So let it be granted that for any two time spans, the assertion
that they are equal to one another has a precise sense grounded in the
intuition of time. Measurement is thereby made possible. And we are
enabled to construct a mathematical theory of time on the foundation we
have indicated: i.e., the basic category “time-point,” the binary relation
253
“A is earlier than B” and the quaternary relation “AB is equal to A
B
.”
Let us now try to express the earlier intuitive sentence in the terms of
these two axioms. We would then have to replace the expression “during
a certain period I saw the pencil lying before me” by “in every time-point
which falls within a certain time span OE.. . .” But such a sentence no
longer reproduces what is intuitively present. Furthermore, it would have
to be admitted as true that,
If Q is a time-point, then the domain of rational numbers to which λ
belongs if and only if there is a time-point P earlier than Q such that
OP = λ · OE can be constructed arithmetically in pure number theory
on the basis of our principles of definition and is thus a real number
in our sense (Ibid., p. 89).
By assuming in this way the time span OE be the axis of the real numbers,
we conclude that to every point Q there thus corresponds a definite real
number as its “abscissa,” and conversely, to every real number there
corresponds a definite time-point. But it is clearly senseless to apply
such a construction when explaining the intuition of time . . . In the first
place, the intuitive continuum is not composed of juxtaposed points.39
Weyl had a similar reflection on the subject of the spatial continuum that
serves as the foundation to the concept of manifold; he shows that it is not
purely and simply reducible to a set of points (see supra). In the second
place, as all our conscious givens are being-now that persist and evolve,
it becomes a flow. We cannot place individual points that are mutually
isolated in this flow.

To every point there corresponds a definite experiential whole; and if


consciousness stands at a certain point, then it has the corresponding
experiential whole; only this is (Ibid., p. 91).

As are the experiences that I may have of this time-point, it is associated

a more or less clear memory whose intentional object is the experi-


ence that I had in a past time-point. [. . .] Thus if I have, say, a visual
perception of brief duration, then, in a moment A, I have not only
this perceptual experience, but simultaneously the memories ‘of’ the
perceptual experiences of all past moments which fall within this brief
period. But not only this; for in this moment A, I remember not only
the perceptual experience in the moment B which occurred a short
time earlier, but the entire experience of this moment B, and this in
turn contains in itself, in addition to the perception, the memories of
the experiences I had in all earlier moments (Ibid).

The similarity with Husserl’s conception is striking. And it is only further


confirmed when, for example, Weyl writes:
254
The view of a flow consisting of points and, therefore, also dissolving
into points turns out to be false. Precisely what eludes us is the nature
of the continuity, the flowing from point to point; in other words, the
secret of how the continually enduring present can continually slip
away into the receding past (Ibid., p. 92).

(iv) The incentives that impel us to move from an intuitive to a mathematical


continuum (separated by a deep chasm) are the same as those

which push science from the experientially constituted reality in which


we live as natural human beings over toward the ‘truly objective’,
exact, non-qualitative, physical world – from the chromatic qualities of
visual things, e.g., to the oscillations of the ether or the corresponding
mathematical descriptions of electromagnetic fields (Ibid., p. 93).

From this point of view, it is justified, according to Weyl, that the only
scientifically viable theory of continuum is one which rests on the mathemat-
ical concept of real numbers, and that, in this sense “the true geometry of
continuity can be developed but only analytically.” (Let us recall that analytic
– or axiomatic – geometry is based on the axiom of continuity according to
which, given a unity span OE, a real number corresponds to every point Q as
abscissa and vice versa.) It would certainly be difficult to contest such a con-
clusion on a strictly mathematical level, although even there one would need
to distinguish several possible meanings of the continuum that are not neces-
sarily equivalent. Furthermore, this does not exclude that the development of
a “truly” philosophy of continuity can all the same be sought.

3. As to the third question, we will only touch here on one of its aspects and
only from a geometric angle. Generally speaking, one can use isomorphic
representation to mathematically express the relation which exists between
objects and the real processes and the images that form themselves via the
organ or the complex of vision in the mental sphere. This notion can be
applied to explain the correspondence between the points of the projec-
tive plane and the qualities of colors. In fact, projective two-dimensional
geometry corresponds to perceived colors and their combinations. More
precisely, colors form a convex region of the complete projective plane
P. The essential property of the convex domain  ⊂ P is that, given two
points A and B, P will contain all the points of the segment included be-
tween A and B at once. Thus, the phenomenon of the perception of colors
can be translated into the following mathematical fact. The quality of per-
ceived color Q can be represented in a continuous manner by the point Q

of the convex region of the projective plane P, Q → Q


, in such a way that
a color Q is produced as a combination of A and B, and represented by a
point Q
located on the segment A
B
; and when we combine A and B with
255
the colors Q and U , the inharmonic ratio of the segments determined on
A
B
by the corresponding points Q
,U
is equal to the inharmonic ratio of
intensities with which A, B(respectively Q, U ) are combined. From this we
conclude that all the theorems of projective geometry applied to the convex
region of the projective plane P allow to an immediate interpretation in
the domain of perceived colors. Let us study this more closely. On the one
hand, we have a manifold of objects V1 – the points of the convex section
of the projective plane – mutually linked by certain relationships such as
R, R
, . . . ; we in particular make the special assumption that the points
are continuously connected, and that point C is situated on the segment
AB. On the other hand, we have a second manifold of objects V2 – that of
perceived colors – characterized by the two following properties:

(i) We can move continuously through the color spectrum from one color
to the next by varying the degree of the principal colors.
(Let us furthermore remember the following facts: (a) the simple qual-
ity of color forms a one-dimensional manifold as it can be determined
by means of a number that varies continuously: the wavelength λ; (b)
that the set of the qualities of combined colors forms, from the phys-
ical point of view, a manifold of infinite dimension; in effect, to fully
describe a combined color, one has to also know the exact type of in-
tensity Jλ by which we can represent each one of the possible (infinite
in principle) wavelengths λ, but these constitute an infinite number
of quantities, namely the independent variables Jλ ; (c) that the col-
ors perceived by our eye form, on the other hand, a two-dimensional
manifold.)
(ii) The color C results from the combination of colors A and B.
If we now make the elements of V2 correspond to the elements of
V1 in such a way that to the elements of V1 , for which the relations
of R, R
, . . . , hold, will correspond the elements of V2 for which
homologous relations hold, we will then say that the both domains of
objects are mutually isomorphic and that, therefore, in this sense, the
two-dimensional projective manifold and the continuous manifold of
colors are isomorphic. Naturally, the foregoing will not have taught us
much about the physical constitution of colors, but it will, nonetheless,
have taught us something fundamental with regard to the geometric
structure of the phenomenal field of colors.

7. Some Philosophical and Mathematical Remarks


in Guise of a Conclusion

In conclusion, we hope that we have helped to show how the results of Husserl’s
research on the constitution of the spatial thing in perception and of objective
256
phenomenal space are of great significance and have been at the origin of
a certain number of important recent developments. Let us add a few more
remarks on this subject. Husserl’s work, in addition to expressing in a rather
precise, albeit intuitive, way a number of subtle mathematical problems, has,
in effect, anticipated numerous fundamental ideas that we have begun to fully
appreciate only very recently. The possible positions that three-dimensional
objects may take in space from the point of view of their perception, as well as
the analysis of their intrinsic structures, are of primary concern here. The prob-
lem lies in the recognition of spatial forms, a central problem in all theories of
perception and one to which neo-Gestalt approaches have greatly contributed.
We can show that these problems can, in principle, express themselves in a
quite rigorous mathematical form and that the concepts of connection and
group will play a fundamental role to this effect. For example, after showing
that

If, in a perception, the series of appearances runs its course in continuous


unity, then the first determination of the change, the so-called differential of
movement, already defines the ‘direction’ of the course, and thereby is given
a system of intentions that are continuously setting out and continuously
getting fulfilled (TS, 1997, p. 86).

And in this way a system of intentional (psycho-physiological) acts that are


engaged and continuously filled is obtained, Husserl notes that the possibility
of the inversion of any direction of change is an essential basis for the process
under consideration. In phenomenological terms, this means that the object
(or body) which exposes itself as a given in the continuity of appearance
remains the same in the inversion of the continuous order, that is to say, when
the temporal course is filled with series of appearances in an exactly opposite
way.

To put it better, the essence of this constitutive presentation includes the


ideal possibility of a reversed presentation in which the same object is
given, and the reversed presentation forms an essential component in the
constitution of full givenness (Ibid, p. 87).

This can be expressed by the following formula: let G be the group of all
rigid movements of object θ j , where j are its different positions, and let σi= j
be the set of rotations of θ , then we have Gσ i (θ j ) = G −1 σi (θ j ).
Husserl very correctly observes that this does not hold for all objects of
perception, and particularly for those that unfold essentially in time, like, for
example, a melody (conceived as a non-static unity, a series of ordered sounds
perceptible as a form), even though this type of object constitutes, much like
those of the spatial order, continuous series or global units. This property of
inversion or of reversibility specifically characterizes objects which we call
257
“spatial bodies.” From a phenomenological point of view, we can affirm that,
whatever the angle (or the point of view) under which we see a rigid object
or, on the other hand, whatever the order in which we make it turn before the
eye, it will not cease appearing like the same object to us. In other words, “[A
spatial object] is what it is only as the identical pole in the systematic unity of
these adumbrations (Abschattungen) or possibilities of adumbration” (Ibid).
In this context, it must be emphasized that “All these continuous series stand
toward one another, and only in the all-encompassing unity of these series
does the [spatial object] come to actual, “omni-sided,” complete givenness [to
perception]” (Ibid).
We can give a simple mathematical interpretation of this phenomenon re-
markably described by Husserl in Ding und Raum. Any spatial object that has
good properties of symmetry can be reconstructed (at least mathematically)
by the set of its invariant properties relative to a given group of rigid transfor-
mations. Elsewhere we show (see Boi, 2004) that it is this kind of hypothesis
that underlies the perception of objects defined by their objective spatial prop-
erties (three dimensionality, among others), and not such as they are perceived
under certain geometric relations of our visual organ. In the more general case
of non-similar objects, we shall say that two among them, say (θ) and (δ), be-
longing to R 3 , are identical or equivalent if they can be continuously deformed
one into the other by a transformation φ of the group ; that is to say, in this
case, the group of Euclidean translations R 3 . The equivalence-classes of (θ),
(δ) will then be obtained by the classification of all the subgroups that are
isomorphic to the principal group  modulo a certain type of transformation
which leave invariant their essential geometric properties.
The idea of the space of perception as a global quotient space made up of
several sensory subspaces – which is not given in advance, but results from
a dynamical and multi-modal process which involves both spatio-temporal
and conscious dimensions and activities – has revealed itself to be very
fruitful for recent research into the relationship between space and percep-
tion. There is, first of all, a philosophical aspect that must be highlighted
before embarking on an analysis of characteristics that are more properly
psycho-physiological. The philosophical aspect is at the heart of the Husser-
lian conception of spatial perception. This last is conceived as a “continuous
formation of the senses,” which amounts to saying that every modality of
sensory perception (of vision, of touch, of hearing, etc.) can be considered
as a perception of the same thing, as an ostension of a spatial figure and
a determined color which overlap and merge. In their unfolding, all these
modalities

function in such a way as to form a sometimes continuous and sometime


discrete synthesis of identification or, better, of unification. This happens
not as a blending of externals; rather, as bearers of ‘sense’ in each phase, as
258
meaning something, the perspectives combine in an advancing enrichment
of meaning and a continuing development of meaning (E. Husserl, The
Crisis of European Sciences, p. 158).

This movement of continuous formation of the sense accompanies each


essential stage of the constitution of the spatial thing and of perception. In
this way, we understand the diversity of partial ostensions of a thing that is one
and the same as forming a system which must be correlated with the diversity
of kinesthetic processes. These, in turn, form a new system which includes
two distinct particular units which pertain to it: internal kinesthetic data (ac-
companied by muscular sensations) and real external body movements. These
are then augmented by a new and important modification: the changing ap-
pearance of being. In the continuity of perception, a spatial body remains
identical when, in the wake of certain kinesthetic sensations, one feels that
these are accompanied by ostensions that belong to them, that is to say, that
the spatial body shows itself in diverse ways while remaining the same. But
an even more significant fact is contained in the inherence of the ostensions of
the spatial body to changing kinesthetic data, namely that this process hides
an intentional active chain of the type “if, then.” In other words,

The exhibitings must occur in a certain systematic order; it is in this way that
they are indicated in advance, in expectation, in the course of a harmonious
perception. The actual kinestheses here lie within the system of kinesthetic
capacity, which is correlated with the system of possible following events
harmoniously belonging to it (Ibid., pp. 161–162).

It is, actually, from this chain of correlations unfolding on an actual and


virtual horizon that the identical spatial thing is constituted. As Husserl un-
derlined, implied in the particular perception of the thing is a whole “horizon”
of non-active and yet co-functioning manners of appearance and syntheses of
validity. Hence, there is not really any place for the singular in perception, as
the perception of a thing always occurs in a perceptual field.
The notion of field is indeed fundamental to the understanding of the
following facts: (1) we always have actual perceptions and a horizon of
possible perceptions; besides, bodily movements, conscious actions and
cognitive acts permit to pass from the former to the latter; (2) three essential
and interrelated processes are involved in perception: the physical one, the
one of perceived phenomena and mental images, and the one of “objects”
of the mind (of thought); (3) finally, the perceptual field always refers us to
a physical field as its substratum, without however being reduced to it.
As Husserl has written:

just as the individual thing in perception has meaning only through an open
horizon of ‘possible perceptions’, insofar as what is actually perceived
259
‘points’ to a systematic multiplicity of all possible perceptual exhibitings
belonging to it harmoniously, so the thing has yet another horizon: be-
sides this ‘internal horizon’ it has an ‘external horizon’ precisely as a
thing within a field of things; and this points finally to the whole ‘world
as perceptual world’. The thing is one out of the total group of simulta-
neously actually perceived things; but this group is not, for us, for con-
sciousness, the world; rather, the world exhibits itself in it; such a group,
as the momentary field of perception, always has the character for us of
a sector ‘of’ the world, of the universe of things for possible perceptions
(Ibid., p. 162).

Hermann Weyl has explained it all in a very eloquent formula: “Our visual
sense evidently does not give us full-value reality, the thing as it really is”
(Weyl, 1934, p. 3).
And further:

The optical image of an object does not simply detach itself from the object
and walk into my brain or present itself to my perception unchanged and
without intermediary; but the light affecting my sense organ is produced by
the object and propagates itself through space according to physical laws.
Therefore the image seen by me by no means renders the object itself but
depends not only on this object but also on all accompanying physical cir-
cumstances. For example, a reflecting surface placed in the path of the light
causes the rays to reach my eye in a manner in which they could reach it
without such an artificial contrivance only if the object of which the reflected
image gives the illusion really existed. In a similar fashion, the refraction
of the sunlight in the water drops of the clouds illusorily displays before my
eyes the rainbow. [. . .] In addition to the influence which the accompanying
physical circumstances exert on our sense perceptions, there is an influence
which penetrates much more deeply and destroys every similarity between
the original and the image, that, namely, of our psycho-physical organi-
zation. The retina [. . .] suppresses by far the greatest part of the ‘lights’
radiating through the universe. But it distorts and reduces in an excessive
manner even the manifold of those which it exhibits (Ibid., pp. 4–7).

A relevant example of this discrepancy between physical reality and per-


ceived phenomenal world is provided by the disparity between the “abun-
dance” of the scale of physical colors and the “paucity” of the colors that are
visually perceived. This is explained by the fact that several colors which are
physically distinct release the same process on the retina and thus physically
produce the same sensation of color. The meaning of this can be shown by the
following analogy. When we make a parallel projection of space on a plane,
all the points that are located in the radius of the projection will coincide in
a single point on the plane; an analogous phenomenon occurs with respect to
the case of different physical colors; they, too, tend to coincide in the same
point on the retina. Therefore, even though phenomenal space is constituted
differently with regard to physical space, it nevertheless shares a certain type
260
of resemblance and geometric similarity with the latter that is fundamental
in every respect to the development of a deeper intelligence of the different
structures of space and of the phenomena which occur in it.

Notes

* Parts of the present work were presented in the previous years at various seminars and
workshops, namely at the École Normale Supérieure and the Collège de France in Paris,
and at the Université Laval and the UQAM in Québec and in Montreal respectively. I am
grateful to Giuseppe Longo, Alain Berthoz, Pierre Kerszberg, Andrew Quinn and Jean-Guy
Meunier for their stimulating and helpful conversations, comments and criticism.
1. Following the usual definition, a subset K of a space X is a knot if K is homeomorphic
with a sphere S p . More generally, K is a link if K is homeomorphic with a disjoint union
S p1 ∪ · · · ∪ S pr of one or more spheres. Two knots or links K , K
are equivalent if there
is a homeomorphism h : X → X such that h(K ) = K
; in other words (X , K ) = (X , K
).
2. For a recent scientific discussion of this important topic, see V. Bruce, P.R. Green, M.A.
Georgeson, Visual Perception. Physiology, Psychology, and Ecology, 3rd edn., especially
chapters 8 and 7 (East Sussex: Psychology Press, 1996).
3. Cf. R. Thom, Apologie du logos (Paris: Hachette, 1990), pp. 163–165; H. Blum, “Biological
Shape and the Visual Science”, Journal of Theoretical Biology 38 (1973): 205–287.
4. The renown Euclid had already given the postulate expressing the law of angular vision
or of the retinal dimension in his Optic (circa 300 B.C.): “Objects that are seen in a wider
angle appear to be larger in dimension from those that are seen in a smaller angle, and
those seen in the same angle appear equal.”
5. Jacques Bouveresse has dedicated to this subject the better part of his detailed and
interesting recent book, Langage, perception et réalité, Éditions Jacqueline Chambon,
Nı̂mes, 1995.
6. Cf. H. Weyl, Mathematische Analyse des Raumproblems (Berlin: Springer, 1923).
7. The introduction in the last century of the concept of manifold by the great German mathe-
matician Bernhard Riemann represents one of the landmark events in the history of mathe-
matics and of knowledge in general. This concept, which led to a complete overhaul, would
prove of extraordinary importance for mathematics. It would not be long before it would
reveal its great value for physics with the theories of special and later general relativity, as
well as for certain branches of the natural sciences, notably chemistry, biology, and botany.
The concept of manifold effectively made possible a first classification of all individuals
and (known) existing elements according to a small number of genera and species based on
their homologous properties, their structural analogies and their common functional affini-
ties. Its true value for the study and the explanation of phenomena in these domains aside,
nothing, in principle, prevents its application for the sake of reaching theoretical models
and developing a certain number of hypotheses. Let M ∞ be a space (a manifold) of infinite
dimensions, in which one seeks to define all the individuals of given natural genus E. Let
V1 , . . . , Vr be a set of subspaces belonging to M, and S1 , . . . , S p the species belonging to
E. Now, let G be the group that acts on M, and H1 , . . . , H p the family of subgroups that
acts respectively on V1 , . . . ,Vr . The set of automorphisms of G that acts transitively on M,
form the global quotient space  = G/M. Let the space M be homogeneous (and possibly
symmetrical), and let it further suffice that the group of transformations G of M be transi-
tive over M. We will then say that any set H1 , . . . , H p (resp. 1 , . . . , p ) can be obtained
from G (or E) by a transformation π of G that leaves the topological structure of M
invariant.
261
8. B. Riemann, “Über die Hypothesen, welche der Geometrie zu Grunde liegen” (Habili-
tationsarbeit, 1854), Abh. Kön. Ges. Wiss., Bd. XIII, 1867; we cite the new edition (see
references).
9. Cf. C. Stumpf, Erkenntnislehre, vol. I, Verlag von J. Ambrosius Barth, Leipzig, 1939, in
particular Part I, § 16.
10. Cf. H. Poincaré, “On the Foundations of Geometry,” The Monist IX(1) (1898): 2–6.
11. Cf. E. Mach, Erkenntnis und Irrtum. Skizzen zur Psychologie der Forschung (Leipzig:
Verlag von J. Ambrosius Barth, 1906).
12. For an analysis in terms of mathematical models of the structure of the organ of hearing,
see C.S. Peskin, “Lectures on Mathematical Aspects of Physiology”, in F.C. Hoppensteadt
(ed.), Lectures in Applied Mathematics, vol. 19 (Providence, Rhode Island: American
Mathematical Society, 1981).
13. Cf. J.F. Herbart, Physiologie als Wissenschaft. Neu gegründet auf Erfahrung, Metaphysik
und Mathematik, Königsberg, 1825.
14. For more recent investigations and interesting results on this subject see, for example, A.
Berthoz, “Reference Frames for the Perception and Control of Movement,” in J. Paillard
(ed.), Brain and Space (Oxford: Oxford University Press, 1991), pp. 81–111. According
to Berthoz, “In order to build a coherent representation of the movement of limbs or body
in space, the brain must assemble the various messages from its sensors so as to select
them according to the environment and to the task in which the subject is engaged, and
must compare them with the expected values generated internally. An essential property
of the processes underlying the construction of a coherent representation of movement
is that it implies multimodal extraction of the important components of the movement
(direction, position, velocity, acceleration, etc.). [. . . ] For example, head motion in space
is reconstructed from visual, vestibular, and somatosensory information. . . All these
operations require reference frames. [. . . ] References frames can be formed in the physical
world, in the sensors themselves, or in the effector system. They do not need to be ‘frames’;
they can be centers of rotation, for instance. They could also be virtual in the sense of
being constructed internally by the brain to perform computations in a topological space,
for instance, in which relative positions or motions are the variables that are processed.
[. . . ] Some properties of the sensory system play a role as reference frames for spatial
orientation, movement, and posture control in humans. It should be stressed that only
vision and the vestibular system can contribute to the measure of head movement in space
during complex movements in which a stable base is absent. They do this by co-operative
mechanism that implies visual–vestibular interaction at many levels of the brain”
(pp. 81–82).
15. W.K. Clifford, “The Postulate of the Science of Space,” in Lectures and Essays, vol. I
(London: Macmillan, 1879), pp. 257–261.
16. Here the word asymmetrical simply refers to objects whose faces have been marked by
different signs, by numbers, for example.
17. See on this subject Piet Hut and Roger N. Shepard, “Turning ‘The Hard Problem’ Upside
Down and Sideways,” Journal of Consciousness Studies 3(4) (1996): 313–329. The authors
of this interesting papers expressed on these hard and delicate issues the following chal-
lenging point of view: “For example, we could consider matter and consciousness both as
emergent properties of underlying and more fundamental aspects of reality. Just as matter
might ultimately be explained as a property of space and time (as is already the case for the
mass and energy of a black hole), so consciousness might be a property of another aspect
of reality, X for short.” And afterwards: “ How reasonable is it, to view matter and mind on
the same level, as complementary? Starting with matter, let us imagine that a future form
of physics will have succeeded in describing matter and energy as forms of excitations
262
of space–time (in analogy with the case of a black hole, say, where space–time curvature
directly provides a definite mass, with no need for any specific ‘matter’ ingredient). Is it,
then, reasonable to turn the hard problem sideways, viewing conscious experience to be
complementary to space–time, neither of the two being reducible to the other?” (p. 322).
18. The German word used by Husserl is Erfüllungen.
19. See D. Willard, “Knowledge,” in B. Smith and D.W. Smith (eds.), The Cambridge
Companion to Husserl (Cambridge: Cambridge University Press, 1995), pp. 154–155.
20. E. Husserl, Logical Investigations (translation of the 2nd edition by J. N. Findlay,
Routledge & Kegan Paul, 1973, p. 724).
21. See E. Husserl, Ding und Raum (Thing and Space), 1907, § 42 to § 44. For interesting
comments on this topic, see K. Mulligan, “Perception,” in B. Smith and D.W. Smith (eds.),
The Cambridge Companion to Husserl (Cambridge: Cambridge University Press, 1995),
pp. 93–97.
22. W. d’Arcy Thompson, Growth and Form (Cambridge: Cambridge University Press, 1917).
23. See B. Smith and D.W. Smith, “Introduction,” The Cambridge Companion to Husserl
(Cambridge: Cambridge University Press, 1995), pp. 20–27.
24. Cf. H. Helmholtz, Die Tatsache in der Wahrnehmung (Berlin, 1878); English translation
in Hermann Von Helmholtz Epistemological Writings, The Paul Hertz/Moritz Schlick
Centenary edition of 1921, edited, with an introduction and bibliography, by R.S. Cohen
and Y. Elkana, Boston Studies in the Philosophy of Science, vol. 37 (Dordrecht: D. Reidel,
1977), pp. 121–122.
25. G. Cantor, “Über unendliche lineare Punktmannigfaltigkeiten. Grundlagen einer
Mannigfaltigkeitslehre,” Mathematische Annalen 21 (1883): 51–58, 545–586.
26. H. Weyl, Die Idee der Riemannschen Fläche (Leipzig: B.G. Teubner, 1913).
27. Cf. I. Strohmeyer (ed.), Studien zur Arithmetik und Geometrie (1886–1901) (The Hague:
Martinus Nijhoff Publishers, 1983) Hua, XXI.
28. E. Husserl, “Geschichtlicher Überblick über die Philosophie der Mathematik,” in op. cit.,
pp. 216–233, Hua XXI.
29. Husserl’s letter to Natorp, by March 14/15, 1897; in B. Bang (ed.), Aufsätze und
Rezensionen (1890–1910) (1979) Hua XXII.
30. See particularly in Thing and Space (English edition of 1997) the chapter 9 on “The
correlation between the visual field and the kinaesthetic sequences” (pp. 139–155), the
chapter 12 on “The typicality of the modifications of appearances in the oculomotor field”
(pp. 191–206), and the chapter 13 on “The constitution of space through the conversion
of the oculomotor field into an expansional and turning manifold” (pp. 207–218).
31. Cf. . “Perception et préhension”, in Apologie du logos, op. cit., p. 173.
32. This seems to follow from the different physiological constitution of the retina’s two
zones. For a complete study of this subject, see M. Imbert, “La vision,” in Ch. Kayser
(ed.), Traité de physiologie, vol. II (Paris: Flammarion, 1976), pp. 1083–1191; and V.
Bruce, P.R. Green, M.A. Georgeson, Visual Perception. Physiology, Psychology, and
Ecology, 3rd edn. (East Sussex: Psychology Press Publishers, 1996), pp. 25–67.
33. Cf. R. Thom, Esquisse d’une Sémiophysique (Paris: InterEditions, 1991), especially
chapter 1; and J. Petitot, Physique du Sens, Éditions du CNRS, Paris, 1992.
34. See, for example, J. Koenderink, Solid Shape (Cambridge, Mass.: MIT Press, 1990).
35. For a rigorous definition of this fundamental mathematical concept, see L.S. Pontryagin,
Topological Groups (New York: Gordon & Brach, 1966).
36. A.V. Shubnikov and V.A. Koptsik, Symmetry in Science and Art (New York: Plenum Press,
1974), particularly chapter 10.
37. Here we refer primarily to the work of 1918, Das Kontinuum; English edition: The
Continuum. A Critical Examination of the Foundation of Analysis (translated by S. Pollard
263
and Th. Bole, Kirksville, MO: Thomas Jefferson University Press, 1987). Hereafter, in
the text we will quote the Dover edition published in 1994.
38. F. Brentano, in S. Körner and R.M. Chisholm, Philosophische Untersuchungen zu Raum,
Zeit und Kontinuum (Hamburg: Meiner, 1976), texts from 1906–1917.
39. See on this subject O. Becker, Beiträge zur phänomenologischen Begründung der Ge-
ometrie und ihrer physikalischen Anwendung, published in Jahrbuch für Philosophische
und Phänomenologische Forschung, edited by Husserl, vol. IV (1923).

References

E.D. Adrian, The Physical Background of Perception (Clarendon Press, Oxford, 1967).
R.B. Angel, “The Geometry of Visibles,” Noûs 8(2) (1974): 87–117.
F. Atteneave, “Criteria for a Tenable Theory of Form Perception,” in W. Wathen-Dunn (ed.),
Models for the Perception of Speech and Visual Form (Cambridge, Mass.: MIT Press, 1967),
pp. 56–67.
A.M. Battro, “Riemannian Geometries of Variable Curvature in Visual Space: Visual Alleys,
Horopters, and Triangles in Big Open Fields,” Perception 5(1976): 9–23.
O. Becker, “Beiträge zur phänomenologischen Begründung der Geometrie und ihrer physikalis-
chen Anwendung,” Jahrbuch für Philosophie und phänomenologische Forschung IV (1923).
A. Berthoz, “Reference Frames for the Perception and Control of Movement,” in J. Paillard
(ed.), Brain and Space (Oxford: Oxford University Press, 1991).
A. Berthoz, Le sens du mouvement (Paris: Odile Jacob, 1996).
H. Blum, “Biological Shape and Visual Science,” Journal of Theoretical Biology 38 (1973):
205–287.
L. Boi, “Le concept de variété et la nouvelle géométrie de l’espace dans la pensée de Bernhard
Riemann,” Archives Internationales d’Histoire des Sciences 134(45) (1995): 82–128.
L. Boi, “Les géométries non euclidiennes, le problème philosophique de l’espace et la con-
ception transcendantale; Helmholtz et Kant, les néo-kantiens, Einstein, Poincaré et Mach,”
Kant-Studien 87(3) (1996): 257–289.
L. Boi, “The Concepts of Connection and Group in the Perception Process,” Journal of Math-
ematical Psychology (2004), forthcoimg.
L. Boi, Le problème mathématique de l’espace. Une quête de l’intelligible (Heidelberg, Berlin:
Springer-Verlag, 1995).
L. Boi, “The Role of Intuition and Formal Thinking in Kant, Husserl and in Modern Mathe-
matics and Physics,” Mathesis (2004), forthcoming.
L. Boi, Géométrie et Phénoménologie de la Perception. Modèles mathématiques et in-
terprétations philosophiques de la perception spatiale (Paris, 2004) (forthcoming).
L. Boi and L. Verner, “Sur quelques aspects phénoménologiques, géométriques et esthétiques
de la perception et de la relation entre surface, forme et couleur,” VISIO, Special Issue
“Cultural Cognition and Space Cognition” 6(2–3) (2001): 205–247.
E.G. Boring, “Visual Perception as Invariance,” Psychological Review 59 (1952): 141–148.
J. Bouveresse, Langage, perception et réalité, vol. 1 (Nı̂mes: J. Chambon Éditeur, 1995).
F. Brentano, Untersuchungen zur Sinnespsychologie (Hamburg: Meiner, 1979), first edition,
1907.
F. Brentano, Psychology from an Empirical Standpoint, in O. Kraus (ed.) (English edition
edited by L.L. McAlister, London: Routledge & Kegan Paul, 1973) (first German edition
published in 1874).
L.E.J. Brouwer, “Consciousness, Philosophy, and Mathematics”, in Collected Works, vol. 1:
Philosophy and Foundations of Mathematics (Amsterdam: North-Holland, 1975).
264
L.E.J. Brouwer, “Über die Abbildung von Mannigfaltigkeiten,” Mathematische Annalen 71
(1911): 97–115.
V. Bruce, P. Green and M. Georgeson, Visual Perception. Physiology, Psychology, and Ecology,
3rd edn. (Sussex: Psychology Press, 1996).
E.H. Carlton and R.N. Shepard, “Psychologically Simple Motions as Geosedic Paths. I. Asym-
metric Objects,” Journal of Mathematical Psychology 34 (1990): 127–188.
É. Cartan, “La géométrie des groupes de transformations,” Journal de Mathématiques Pures
et Appliquées 6 (1927): 11–119.
E.C. Carterette and M.P. Friedman, Handbook of Perception, vol. III: Biology of Perceptual
Systems (New York: Academic Press, 1974).
E. Cassirer, “The concept of Group and the Theory of Perception,” Philosophy and Phenomeno-
logical Research IV (1944).
M. Cesa-Bianchi, A. Beretta and R. Luccio, La percezione (Milan: Franco Angeli Editore,
1977).
K.C. Chen, “A Group-Theoretic Analysis of Mental Space Isometries in Form Recognition,”
Journal of Mathematical Psychology 37(1993): 401–420.
U. Claesges, Edmund Husserls Theorie der Raumkonstitution (The Hague: Martinus Nijhoff,
1964).
W.K. Clifford, Lectures and Essays, vol. I (London: Macmillan, 1879).
H. Conrad-Martius, “Zur Ontologie und Erscheinungslehre der realen Außenwelt,” Jahrbuch
für Philosophie und phänomenologische Forschung 3(1920): 345–396.
Da Silva, “Husserl’s Phenomenology and Weyl’s Predicativism,” Synthese 110(2) (1997): 277–
296.
M. Delbrück, Mind from Matter? (Palo Alto, California: Blackwell Scientific Publishers, 1986).
Ch. Ehrenfels, “Über Gestaltqualitäten,” Vierteljahrschrift für wissenschaftliche Philosophie
XIV (1890): 249–292.
W.D. Ellis (ed.), A Source Book of Gestalt Psychology (London: Routledge & Kegan Paul,
1967).
F. Enriques, “Sulla spiegazione psicologica dei postulati della geometria,” Rivista di Filosofia
IV(3) (1901): 171–195.
D. FØllesdal, “Phenomenology,” in E.C. Carterette and M.P. Friedmann (eds.), Handbook of
Perception (New York: Academic Press, 1974), pp. 377–386.
J.J. Gibson, The Perception of the Visual World (Boston: Houghton Mifflin, 1950).
J.J. Gibson, The Senses Considered as Perceptual System (Boston: Houghton Mifflin,
1966).
R.L. Gregory, The Intelligent Eye (London: Weidenfeld and Nicolson, 1970).
R.L. Gregory, Eye and Brain. The Psychology of Seeing, 5th edn. (Princeton: Princeton Uni-
versity Press, 1997).
N.R. Hanson, Patterns of Discovery, An Inquiry into the Conceptual Foundations of Science
(Cambridge: Cambridge University Press, 1958).
C.W. Harvey, Husserl’s Phenomenology and the Foundations of Natural Science (Athens: Ohio
University Press, 1989).
S. Helgason, Differential Geometry, Lie Groups and Symmetric Space (New York:
Academic Press, 1978).
H. Helmholtz, Handbuch der physiologischen Optik, 3 vols. (Leipzig: J.A. Barth, 1867).
H. Helmholtz, Wissenschaftliche Abhandlungen, vol. II (Leipzig: J.A. Barth, 1883).
J.F. Herbart, Psychologie als Wissenschaft. Neu gegründet auf Erfahrung, Metaphysik und
Mathematik (Königsberg, J. A. Barth, 1825).
E. Hering, Grundzüge der Lehre vom Lichtsinn (Berlin: Springer-Verlag, 1905).
D. Hilbert and S. Cohn-Vossen, Anschauliche Geometrie (Berlin: Springer-Verlag, 1932).
265
W.C. Hoffman, “The Lie Algebra of Visual Perception,” Journal of Mathematical Psychology
3 (1966): 65–98.
J. Hopking, “Visual Geometry,” The Philosophical Review 82(1) (1973): 3–34.
I.P. Howard, “Orientation and Motion in Space,” in Handbook of Perception, vol. III: Biology
of Perceptual Systems (New York; Academic Press, 1974).
E. Husserl, Logische Untersuchungen, vols. I–III (Tübingen: Max Niemeyer, 1980) (Logical
Investigations, English edition translated by J.N. Findlay, London: Routledge & Kegan Paul,
1973).
E. Husserl, The Crisis of European Sciences and Transcendental Phenomenology: An Introduc-
tion to Phenomenological Philosophy (translated by D. Carr, Evanston, Ill.: Northwestern
University Press, 1970).
E. Husserl, Formal and Transcendental Logic (translated by D. Cairns, The Hague:
Nijhoff, 1969).
E. Husserl, Ding und Raum, Vorlesungen 1907 (herausgegeben von U. Claesges, The Hague:
Martinus Nijhoff, 1973) (english translation by R. Rojcewicz, Thing and Space – Lectures
of 1907, Dordrecht: Kluwer Academic Publishers, 1997).
E. Husserl, in I. Strohmeyer (ed.), Husserliana – Gesammelte Werke, vol. XXI: Studien zur
Arithmetik und Geometrie (The Hague: Nijhoff, 1983), texts from 1886–1901.
E. Husserl, Ideas Pertaining to a Pure Phenomenology and to a Phenomenological Philosophy.
First book: General Introduction to a Pure Phenomenology (translated by F. Kersten, The
Hague: Martinus Nijhoff, 1982).
E. Husserl, Ideas Pertaining to a Pure Phenomenology of Constitution. Second book: Studies in
the Phenomenology of Constitution (translated by R. Rojcewicz and A. Schuwer, Dordrecht:
Kluwer Academic Publishers, 1989).
P. Hut, “Turning ‘The Hard Problem’ Upside Down and Sideways,” Journal of Consciousness
Studies 3(4) (1996): 313–329.
P. Hut, “Structuring Reality: The Role of Limits,” in J.L. Casti and A. Karlqvist (eds.), Bound-
aries and Barriers. On the Limits to Scientific Knowledge (Reading, MA: Addison-Wesley,
1996), pp. 148–187.
M. Imbert, “La vision,” in Ch. Kayser (ed.), Traité de physiologie, tome II (Paris: Flammarion,
1976), pp. 1083–1191.
W. James, “The Perception of Space. (I) and (II),” Mind 45 (1887): 1–30; 47 (1887): 183–
211.
W. James, The Principles of Psychology, vol. I (Cambridge, MA: Harvard University Press,
1981).
G. Kanizsa, Organization in Vision (New York: Praeger, 1979).
I. Kant, Kritik der reinen Vernunft (1781–1787) (Hamburg: Felix Meiner Verlag, 1990).
P. Kerszberg, “Of Exact and Inexact Essences in Modern Physical Science,” in L. Hardy
and L. Embree (eds.), Phenomenology of Natural Science (Amsterdam: Kluwer Academic
Publishers, 1992).
S. Kobayashi, K. Nomizu, Foundations of Differential Geometry, vols. I and II (New York:
Interscience Publishers, 1963).
K. Koffka, Principles of Gestalt Psychology (New York: Harcourt, 1935).
W. Köhler, Die physischen Gestalten in Ruhe und in stationärem Zustand (Erlangen: Verlag
der Philosophischen Akademie, 1924).
W. Köhler, Gestalt Phychology (New York: H. Liveright, 1929).
D. Lohmar, Phänomenologie der Mathematik, in Phænomenologica (Dordrecht: Kluwer Aca-
demic Publishers, 1989).
G. Longo, “The Mathematical Continuum: From Intuition to Logic,” (Paris: Ecole Normale
Supérieure), Preprint, LIENS 97/12, pp. 3–29.
266
Z.K. Lorenz, The Role of Gestalt Perception in Animal and Human Behaviour, in L.L. Whyte
(ed.), Aspects of Form (New York: Elsevier Publishing Company, 1968).
E. Mach, Erkenntnis und Irrtum. Skizzen zur Psychologie der Forschung (Leipzig: J.A. Barth,
1906).
K. Mulligan, “Perception,” in B. Smith and D.W. Smith (eds.), The Cambridge Companion to
Husserl (Cambridge: Cambridge University Press, 1995), pp. 168–238.
F. Nef, L’objet quleconque (Paris: Vrin, 1998).
N. O’Connor and B. Hermelin, Seeing and Hearing and Space and Time (London: Academic
Press, 1978).
J. Paillard, “Motor and Representational Framing of Space,” in J. Paillard (ed.), Brain and Space
(Oxford; Oxford University Press, 1991).
S.E. Palmer, “The Role of Symmetry in Shape Perception,” Acta Psychologica 59 (1985):
67–90.
S.E. Palmer, “Modern Theories of Gestalt Perception,” in G.W. Humphreys (ed.), Understand-
ing vision (Oxford: Blackwell, 1992).
R. Penrose, The Emperor’s New Mind. Concerning Computers, Minds, and The Laws of Physics
(Oxford: Oxford University Press, 1989).
C.S. Peskin, Lectures on Mathematical Aspects of Physiology, in F.C. Hoppensteadt (ed.), Lec-
tures in Applied Mathematics, vol. 19 (Providence, Rhode Island: American Mathematical
Society, 1981).
J. Petitot, “Structuralisme et phénoménologie,” in J. Petitot (ed.), Logos et Théorie des Catas-
trophes, Actes du Colloque de Cerisy en l’honneur de René Thom (Genève: Ed. Patiño,
1989), pp. 345–376.
J. Petitot, “Le physique, le morphologique, le symbolique: remarques sur la vision,” Revue de
Synthèse (1–2) (1990): 139–183.
J. Petitot, Physique du Sens, Editions du CNRS, Paris, 1992.
H. Poincaré, “L’espace et la géométrie,” in La Science et l’Hypothèse, Flammarion, Paris, 1902.
H. Poincaré, “The Foundation of Geometry,” The Monist IX(1) (1898): 1–43.
L.S. Pontryagin, Topological Groups (first Russian edition, 1938, New York: Gordon & Breach,
1966).
N. Rashevsky, “Physico-Mathematical Aspects of the Gestalt Problem,” Philosophy of Science
1 (1934): 409–419.
A.H. Reinhardt-Rutland, “Perceiving the orientation-in-depth of triangular surfaces: static-
monocular, moving-monocular, and static-binocular viewing”, The Journal of General Psy-
chology, Vol. 123, No. 1, 1996, pp. 19–28.
W. Richards, “Visual Space Perception,” in E.C. Cartarette and M.P. Friedman (eds.), Handbook
of Perception, vol. V: Seeing (New York: Academic Press, 1975), pp. 351–386.
B. Riemann, in R. Narasimhan (ed.), Gesammelte mathematische Werke, wissenschaftlicher
Nachlaß und Nachträge/Collected Papers (Berlin: Springer-Verlag, 1990).
I. Rock, Orientation and Form (New York: Academic Press, 1973).
D. Rolfsen, Knots and Links (Berkeley, CA: Publish or Perish Press, 1979).
W. Schapp, Beiträge zur Phänomenologie der Wahrnehmung (Wiesbaden: Heymann, 1976).
R.N. Schepard, “The Circumplex and Related Topological Manifolds in the Study of Percep-
tion,” in S. Shye (ed.), Theory Construction and Data Analysis in the Behavioral Sciences
(San Francisco: Jossey-Bass, 1978).
E. Schrödinger, “Grundlinien einer Theorie der Farbenmetrik im Tagessehen” (1920), in
Gesammelte Abhandlungen, vol. 4 (Wien: Verlag der Österreichischen Akademie der Wis-
senschaften, 1984).
A.V. Shubnikov and V.A. Koptsik, Symmetry in Science and Art (New York: Plenum Press,
1974).
267
C. Sinigaglia, La seduzione dello spazio, Geometria e filosofia nel primo Husserl (Milan:
Edizioni Unicopli, 2000).
B. Smith and D.W. Smith, “Introduction,” in B. Smith and D.W. Smith (eds.), The
Cambridge Companion to Husserl (Cambridge: Cambridge University Press, 1995), pp.
1–44.
G. Sperling, “Mathematical Models of Binocular Vision,” in S. Grossberg (ed.), Mathematical
Psychology and Psychophysiology, SIAM/AMS Proceedings, vol. 13 (Providence: American
Mathematical Society, 1981).
N. Steenrod, The Topology of Fibre Bundles (Princeton: Princeton University Press, 1951).
C. Stumpf, Erekenntnislehre, vol. 1 (Leipzig: Verlag von J.A. Barth, 1939).
C. Stumpf, Über den psychologischen Ursprung der Raumvorstellung (Leipzig, 1873).
P. Suppes, “Is Visual Space Euclidean?,” Synthese 35 (1977): 397–421.
R. Thom, Apologie du logos (Paris: Hachette, 1990).
R. Thom, Esquisse d’une Sémiophysique (Paris: InterEditions, 1991).
D’Arcy Thompson, On Growth and Form (Cambridge: Cambridge University Press, 1917).
W.R. Uttal, Visual Form Detection in 3-Dimensional Space (New Jersey: Lawrence
Erlbaum Associates, 1983).
M. Wagner, “The Metric of Visual Space,” Perception and Psychophysics 38(6) (1985): 483–
495.
R. Walk and H.L. Pick Jr. (eds.), Intersensory Perception and Sensory Integration (New York:
Plenum Press, 1981).
M. Wertheimer, “Untersuchungen zur Lehre von der Gestalt. II,” Psychologische Forschung 4
(1923): 301–350.
H. Weyl, Das Kontinuum. Kritische Untersuchungen über die Grundlagen der Analysis
(Leipzig: Veit, 1918) (The Continuum. A Critical Examination of the Foundation of Analysis
(english edition translated by S. Pollard and Th. Bole, Kirksville, Missouri: The Thomas
Jefferson University Press, 1987), Dover edition, 1994.
H. Weyl, Mathematische Analyse des Raumproblems (Berlin: Springer, 1923).
H. Weyl, Mind and Nature (Philadelphia: University of Pennsylvania Press, 1934).
H. Weyl, Philosophy of Mathematics and Natural Science (Princeton: Princeton University
Press, 1949).
H. Weyl, Symmetry (Princeton: Princeton University Press, 1952).
F.R.S. Zeeman, “Brain Modeling,” in Structural Stability, The Theory of Catastrophes, and
Applications in the Sciences 525 (1976): 367–372.
S. Zeki, A Vision of the Brain (London: Blackwell Scientific Publishers, 1993).
D.W. Zimmerman, “Could Extended Objects be Made Out of Simple Parts? An Argument for
‘Atomless Gunk’,” Philosophy and Phenomenology Research 56(1) (1996): 1–29.
L. Zusne, Visual Perception of Form (New York: Academic Press, 1970).

Potrebbero piacerti anche