Download as pdf or txt
Download as pdf or txt
You are on page 1of 436

LOOKING INTO PICTURES

This page intentionally left blank


LOOKING INTO PICTURES
AN INTERDISCIPLINARY APPROACH TO PICTORIAL SPACE

edited by
Heiko Hecht
Robert Schwartz
Margaret Atherton

A Bradford Book
The MIT Press
Cambridge, Massachusetts
London, England
© 2003 Massachusetts Institute of Technology

All rights reserved. No part of this book may be reproduced in any form by any
electronic or mechanical means (including photocopying, recording, or informa-
tion storage and retrieval) without permission in writing from the publisher.

This book was set in Times New Roman on Quark by Asco Typesetters, Hong
Kong.
Printed and bound in the United States of America.

Library of Congress Cataloging-in-Publication Data

Looking into pictures : an interdisciplinary approach to pictorial space / edited by


Heiko Hecht, Robert Schwartz, Margaret Atherton.
p. cm.
“A Bradford book.”
Includes bibliographical references and index.
ISBN 0-262-08310-8 (alk. paper)
1. Picture perception—Congresses. 2. Perspective—Congresses. 3. Picture
interpretation—Congresses. I. Hecht, Heiko. II. Schwartz, Robert, 1940–
III. Atherton, Margaret.
BF243 .L66 2003
152.14—dc21 2002070319

10 9 8 7 6 5 4 3 2 1
Contents

Acknowledgments ix
Introduction xi

PART I
THE DUAL NATURE OF PICTURE PERCEPTION 1

Chapter 1
In Defense of Seeing-In 3
Richard Wollheim

Chapter 2
Conjoint Representations and the Mental Capacity for Multiple Simultaneous
Perspectives 17
Rainer Mausfeld

Chapter 3
Relating Direct and Indirect Perception of Spatial Layout 61
H. A. Sedgwick

Chapter 4
The Dual Nature of Picture Perception: A Challenge to Current General Accounts
of Visual Perception 77
Reinhard Niederée and Dieter Heyer

Chapter 5
Perceptual Strategies and Pictorial Content 99
Mark Rollins
vi Contents

PART II
THE STATUS OF PERSPECTIVE 123

Chapter 6
Optical Laws or Symbolic Rules? The Dual Nature of Pictorial Systems 125
John Willats

Chapter 7
Perspective, Convention, and Compromise 145
Robert Hopkins

Chapter 8
Resemblance Reconceived 167
Klaus Sachs-Hombach

Chapter 9
What You See Is What You Get: The Problems of Linear Perspective 179
Klaus Rehkämper

Chapter 10
Pictures of Perspective: Theory or Therapy? 191
Patrick Maynard

PART III
THE NATURE AND STRUCTURE OF RECONCEIVED PICTORIAL
SPACE 213

Chapter 11
Reconceiving Perceptual Space 215
James E. Cutting

Chapter 12
Pictorial Space 239
Jan J. Koenderink and Andrea J. van Doorn

Chapter 13
Truth and Meaning in Pictorial Space 301
Sheena Rogers
vii Contents

Chapter 14
Line and Borders of Surfaces: Grouping and Foreshortening 321
John M. Kennedy, Igor Juricevic, and Juan Bai

Chapter 15
Irreconcilable Views 355
Hermann Kalkofen

References 379
Contributors 405
Index 407
This page intentionally left blank
Acknowledgments

The last half of the twentieth century witnessed revolutionary changes in


the study of perception. The assumptions underlying the earlier percep-
tion paradigm—(1) the eye as camera, (2) the visual image as a picture,
and (3) perception as primarily the result of interpreting static pictorial
cues—have all been severely challenged. The implications of this para-
digm shift for the study of picture perception, however, have not been well
explored.
The editors thought the time was ripe for an in-depth, interdisciplinary
examination of this issue. Aided by a grant from the Zentrum für inter-
disziplinäre Forschung (Center for Interdisciplinary Research; ZiF) in
Bielefeld, Germany, we contacted many of the most prominent philoso-
phers and psychologists in Europe and North America who are actively
engaged in research on pictorial representation and picture perception.
We asked them each to write a chapter on the topic “reconceiving picto-
rial space” (as described in the introduction) and to participate in a con-
ference to discuss their work. Just about everyone contacted agreed on the
need and importance of the project and also agreed to write a paper. Sub-
sequently, chapter drafts of the participants were circulated, and a week-
long meeting was convened at the ZiF in the summer of 2000 to discuss
and debate each person’s position.
We thank the staff at the ZiF for their excellent administrative support.
Johannes Roggenhofer, director of the ZiF, as well as Michael Kubovy
and William Epstein participated in the conference discussions, and with-
out their input the book would not be the same.
This page intentionally left blank
Introduction

Once Kepler correctly characterized how light reflected from objects pro-
jects a pattern of rays on the retina, a close relation between pictorial space
and perceived physical space became a guiding presupposition of most
work in the theory of vision. Given that the stimulus for vision is an
image, a picture of the world the light rays “paint” on the retina, percep-
tion of the spatial layout must somehow depend on processing this image.
We see the world through, or on the basis of, “picture images.” The link
between pictorial and physical space became even tighter as a result of the
Renaissance development of systems for perspective rendering. Pictures
constructed according to the rules of perspective seemed to mirror, and
hence resemble, what they depicted. In the case of trompe l’oeil, viewers
could not even tell the difference between the two. Perspective pictures
were thought, therefore, to provide a correct, perhaps the only correct,
way to represent physical space pictorially.
If any of this was in doubt, the subsequent invention of the camera
clinched the case for allying pictorial and perceived space. Photographs,
it was widely thought, represent space as it “really” is. Moreover, the
similarities between the eye and the camera are striking. As the camera
focuses light through a pinhole and lens onto a flat piece of film, so the
retinal image is a projection of the world through a lens and pinhole (iris)
onto the two-dimensional retinal surface. In both cases the projections
follow the path of linear perspective. Thus, the “eye as camera” metaphor
was born, and it came to dominate and set the problematic for the study
of perception. As influentially formulated by Hermann von Helmholtz
(1894) in the late nineteenth century, the goal of perceptual psychology
was to find the rules of inference our visual system uses to derive a unique
3-D interpretation from the 2-D retinal image.
xii Introduction

By the middle of the twentieth century, many of the assumptions under-


lying the “eye as camera” metaphor, however, came under serious attack.
Philosophers questioned the very coherence of the notion of a sensory
“picture,” an uninterpreted phenomenal entity, standing as an interme-
diary between our perceptions and the physical world. Meanwhile, in the
psychology of vision, theorists, building on the pioneering work of James
J. Gibson (1950, 1966, 1979), began to propose models of spatial percep-
tion at odds with the whole conception of vision as a process of inference
from a static pictorial image.
Amidst all these changes, however, one fundamental assumption has
continued to play a prominent role in both the philosophical and psycho-
logical studies of picture perception. The idea that perspective pictures are
special, the true way to represent space, although periodically challenged,
has not been displaced. When philosophers talk about pictures and how
pictures differ from languages and other forms of representation, it is
almost always photographs or perspective renderings they have in mind.
Similarly, in psychology, the study of picture perception tends to center on
phenomena largely concerned with the viewing of perspective pictures.
The special status accorded perspective pictures in vision theory is further
highlighted when one considers that many experiments aimed at studying
nonpictorial spatial perception are not actually conducted in real physical
environments. Subjects, instead, are provided perspective pictures, or their
computer screen equivalents, to view.
By contrast, theory and practice in artistic circles have evolved along a
rather different path. Although the influence of perspective doctrines on
post-Renaissance Western art was profound, artists eventually felt them-
selves needlessly constrained by its demands. Twentieth-century Western
art has seen a veritable explosion of alternative schemes for depicting
space. Today, the idea of there being only one legitimate way to represent
space in artistic renderings, the one the post-Renaissance tradition
favored, seems old-fashioned if not, as some contend, an outright matter
of cultural bias or chauvinism. Surprisingly, philosophical and psycho-
logical work on picture perception has been relatively unmoved by these
shifts in artistic practice and attitude. But the rationale for this steadfast
adherence to traditional commitments is open to question.
As noted, the supporting “eye as camera” metaphor that has served to
underpin the paradigm is no longer sacrosanct. In addition, other argu-
ments commonly used to justify according perspective pictures special
status, have also come under attack. For example, a number of influential
xiii Introduction

writers in psychology and philosophy have argued forcefully that appeals


to the notion of “resemblance” fail to explain either the character or
uniqueness of perspective representation. Some theorists, in fact, have
gone on to argue that perspective representation itself depends on con-
ventions and so is best understood as a form of symbolization, on par with
languages, music notation, maps, graphs, and the like.
All volume contributors were asked to consider what, if anything, recent
developments in their area of specialty entailed about the relationship
between pictorial and perceived physical space. More particularly, they
were asked to consider such questions as these: Is the traditional con-
ception of the nature of pictorial space still adequate, or must it be recon-
ceived? Are vision scientists justified in substituting perspective pictures
for real objects and environments in their studies? Are these setups, in
Gibson’s terms, “ecologically valid”? Given the pervasiveness of non-
perspective renderings, not only in art but in advertising, scenic design,
instructional displays, and so many other areas, does it make sense to treat
these forms of picturing as of only marginal interest? Might a reconcep-
tion of pictorial space, then, enable vision science to contribute more to
our understanding of this wider realm of pictorial representations? And
what might the implications of such a reorientation be for work on com-
puter-generated “virtual reality” environments?
Finally, any reconception of “pictorial space” is sure to have profound
implications for the conduct of experimental research in picture percep-
tion. For the assumptions of the traditional paradigm not only have
shaped the theories developed; they have largely determined the very
problems considered suitable or important to study. For example, much
contemporary work in picture perception is devoted to testing and
accounting for the effects that alterations in viewing angle, height, motion,
and distance have on the way we see pictures. These experiments, how-
ever, tend to hinge crucially on features of perspective pictures and the
geometry of the light rays such pictures project. Accordingly, the ratio-
nale for the studies, as well as the explanations proffered, depend on
assumptions about the relationship between perspective renderings and
perceived space. But if the account of pictorial space and pictorial repre-
sentation were to change, what types of experimental methods and mea-
surement devices would be most suitable to explore the domain?
The papers of the contributors fall into three natural groups: the dual
nature of picture perception, the status of perspective, and exploring pic-
torial space.
xiv Introduction

I THE DUAL NATURE OF PICTURE PERCEPTION

In addition to representing people, places, and things, pictures themselves


are physical objects, with their own size, shape, color, and location. For
the most part observers are well aware of these properties. Thus, the
perception of pictures appears to be different in important ways from
the perception of ordinary objects and spatial layouts. Any satisfactory
theory of pictorial space must take into account this dual nature of picture
perception. And any reconception of pictorial space must do so as well.
The chapters in part I offer competing views on this distinctive aspect of
picture perception.
In chapter 1, Richard Wollheim elaborates his influential thesis that
pictorial representation is to be understood in terms of a phenomenal
notion of “seeing-in.” He claims that this account provides a natural
framework for handling the dual nature (or “twofoldness”) of picture per-
ception. Wollheim also argues that his seeing-in approach better explains
pictorial representation than other prominent theories do, in particular
inference and illusion models. For Wollheim, representational content is
set by the intentions of the artist and revealed to observers by what they
are able to “see in” the picture.
Rainer Mausfeld, in chapter 2, holds that picture perception does not
present a separate or special explanatory problem. Rather, the duality of
picture perception is symptomatic of the perceptual system’s wider capac-
ity to handle what Mausfeld terms “conjoint representations.” Building
on a basic theoretical model in which sensory inputs compete to trigger
representational primitives of the perceptual system, Mausfeld proposes
that the duality of picture perception is one of a number of cases that can
be explained by hypothesizing that more than one primitive is triggered.
Mausfeld discusses many other cases that might exemplify a similar mech-
anism, for example, color and brightness, which, as primitives result in
the simultaneous perception of surface illumination and ambient illumi-
nation. Mausfeld concludes by using the theoretical apparatus he has set
up to give a new interpretation of Irvin Rock’s “proximal mode.”
In chapter 3, H. A. Sedgwick links the dual nature of pictures with the
fundamental distinction between direct and indirect perception. Building
on Gibson’s ideas, he argues that indirect perception is reducible to direct
perception. But Sedgwick allows that at present the latter theory is far
from adequate. The dual nature of pictures will remain problematic, there-
fore, until a better account of direct perception is at hand. Summarizing a
xv Introduction

series of Gibsonian-inspired experiments, Sedgwick provides striking evi-


dence that “cross-talk” between directly and indirectly specified aspects
of pictures can explain most of the “errors” found in recovering physical
space from pictures.
Reinhard Niederée and Dieter Heyer argue, in chapter 4, that we should
not take the dual aspects of picture perception as a kind of two-stage
process. Instead, the duality reflects two aspects of a single unified per-
ceptual process, the nature of which has implications beyond the case of
pictures. Their proposal is that the duality of picture perception is an indi-
cation of the visual system’s capacity to manage ambiguity. This sort of
ambiguity, they claim, is found not only in picture perception but also in
ordinary vision. Here, too, ambiguity is frequently preserved rather than
reduced to a single consistent scene. Using concepts such as Sedgwick’s
cross-talk, they demonstrate the functioning of interactive processes that
require the presence of both elements and conclude that perception can
profitably be considered a complex phenomenon, more like a language
than a simple presentation of what is out there.
In chapter 5, Mark Rollins begins by explaining why strategic design
theory (SDT) is taken by many to provide a serious alternative to Marr-
style computational models of vision. This “bag of tricks” account appeals
to a variety of flexible, context-sensitive, task-specific operations that do
not require highly articulated representational structures. Rollins explores
the implications that adoption of such an SDT approach has for a range
of issues in picture perception, in particular the issues related to direct
versus indirect perception. Among the other topics he examines are the
issues of conceptual versus nonconceptual content, and recognitional
(seeing-in) versus conventionalist accounts of representation.

II THE STATUS OF PERSPECTIVE

The dominant view of pictorial space gives pride of place to perspective


renderings. But given the dual nature of picture perception, it cannot be
claimed that seeing even the most accurately constructed perspective pic-
ture is the same as seeing that which it represents. In addition, we know
that people on an everyday basis deal perfectly well with a myriad of pic-
tures that violate the canons of perspective. So wherein lies the special
status of this system of depiction? A common answer is that these render-
ings provide a true or realistic rendering of space. But this response will
not be satisfactory unless it is backed up by a plausible analysis of the idea
xvi Introduction

of one representation of pictorial space being correct or more accurate


than another. The chapters in part II discuss the justification for claims
about the special status of perspective.
Chapter 6, John Willats’s chapter, serves as a bridge between parts I
and II. Willats reprises the old question of whether the rules of pictorial
representation are mere convention or immutable natural law. His answer
is that the duality of pictures is mainly an outcome of the method used to
describe the representational system. In this way, he seeks to relate the
phenomenal duality of picture perception with the supposed unique status
of perspective renderings. One outcome of his analysis is an explanation
for why line drawings, which violate almost every rule of perspective, look
as convincing as they do.
Robert Hopkins argues in chapter 7 that the question of whether per-
spective is conventional is itself unclear, and he offers a more precise
statement of the issue. Hopkins then presents his own account of picto-
rial representation in terms of “seeing-in” as experienced resemblance of
visual form. One major challenge resemblance theories face is that they
have trouble accounting for the wide range of pictorial renderings that do
not seem to resemble what they represent. Hopkins counters this objection
by loosening the seeing-in criterion. This move also provides room for
conventional factors to play a role in assessing resemblance. Nevertheless,
Hopkins concludes that these compromises do not undercut his central
claim that perspective picturing is special and nonconventional.
In chapter 8, Klaus Sachs-Hombach sets out to develop a principle for
distinguishing pictures from linguistic and other symbolic forms. He, too,
rejects a conventionalist theory in favor of a resemblance-based account
of pictorial representation. The principle that evolves from his study is “A
sign is a picture if the perception of the essential properties that the sign
vehicle has in relevant respects is identical to the perception one would
have of the corresponding properties of some other object under a certain
perspective and if this perception is constitutive for the interpretation of
the sign.” The chapter is an attempt to explain and justify the ideas pro-
posed and appealed to in this definition.
Klaus Rehkämper in chapter 9 defends the view that linear perspective
can be identified as the correct way to represent the world pictorially,
because it captures the way the visual system sees the world. Like Sachs-
Hombach, he is concerned with addressing problems raised by Nelson
Goodman’s claim that linear perspective represents conventionally rather
than representing by some natural relation. Rehkämper suggests that
xvii Introduction

Goodman’s argument embodies a confusion of the relationship between


the picture and its object with the relationship between the picture and an
observer of that picture.
In chapter 10, Patrick Maynard illustrates the consequences of what he
identifies as a fallacy: the fallacy that a linear projection is a depiction and
hence that a theory of linear projection is a theory of depiction. He urges
the theoretical importance of recognizing that linear perspective is no more
than a technological tool that can be used to make pictures. Moreover, as
essentially a matter of geometry, the tool has precisely nothing to say
about vision or subjectivity. Maynard urges development of an indepen-
dent theory of depiction, one that recognizes we have a more complicated
task and even a more complicated spatial task than merely relating some-
thing two-dimensional to something three-dimensional.

III EXPLORING PICTORIAL SPACE

However one settles the theoretical issues in parts I and II, there remains
the practical problem of applying the answers in the conduct of experi-
mental research. Alternative conceptions of pictorial representation and
pictorial space structure the problem space, the evidence sought, and the
methods of collecting, measuring, and interpreting the data. Just how
these matters play out will, of course, vary somewhat with the topic of
concern (e.g., size, shape, and distance recognition). The chapters in part
III explore several core empirical issues against the background of con-
cerns discussed in the previous parts.
James Cutting’s chapter, chapter 11, provides a transition between parts
II and III. Cutting proposes solutions to a variety of theoretical questions
about perception and perspective. He suggests that all perceived space has
an inherent ordinal metric, thereby reuniting the dual aspects of pictures
and creating an elegant synthesis of direct and indirect perception. By
distinguishing between close personal space and a farther vista space,
Cutting puts pictures in their (vista) place, which amounts to a perceiver
mode where would-be spatial distortions are no longer an issue. He also
offers us a new explanation of why line drawings work as well as they do.
In chapter 12, Jan Koenderink and Andrea van Doorn offer an in-depth
analysis of the problem of measuring pictorial space, exploring along the
way why common alternative attempts to measure a space that is non-
physical have gone wrong. Their own nonreductionist approach is uncon-
ventional and constructive. Koenderink and van Doorn first discuss their
xviii Introduction

development and use of a number of methods to access observers’ im-


plicit knowledge of pictorial space. Then, on the basis of evidence from
their empirical studies of surface relief, they derive a formal description
of pictorial space as distinctly different from Euclidean space.
Sheena Rogers, in chapter 13, puts forth and defends the thesis that a
realist theory of picture perception is problematic. Relying on ideas asso-
ciated with ecological approaches to vision, she questions the doctrine
that if the same information is available in pictures as is in a real scene the
picture must look real. Moreover, she provides a body of empirical data
showing that this need not be so. Similarity does not suffice for realism.
Rogers also offers an explanation of why pictures may “lie” to us and why
trompe l’oeil is so powerful.
In chapter 14, John Kennedy, Igor Juricevic, and Juan Bai set out to
solve a mystery that can be called the “aggravated underspecification
problem.” A full-fledged rendition, such as a photograph, already under-
specifies the rendered scene because the picture is necessarily less complex.
Simple line drawings substantially widen this gap. How is it, then, that we
readily see faces, objects, and so on, in a few marks on a piece of paper or
even in one continuous pencil line? The visual system apparently finds
an empirical solution to a problem that has no theoretical solution. The
authors attempt to solve this mystery by studying the significance of lines
and borders in pictures. They show that the empirical solution is found, in
part, by interpreting borders as edges of surfaces. The ecological view
reflected in Kennedy and his coauthors’ argument supplements Koen-
derink and van Doorn’s attempt to use image statistics as an explanation
for this mysterious ability of the visual system. It also contrasts with
Willats and Cutting’s solutions to the same problem.
Hermann Kalkofen also provides reasons to believe there is no easy
relationship between pictorial space and physical space. He guides us
through a number of neglected examples of pictorial representations that
do not presume one point of view at one time. These representations
embody numerous inconsistencies, and yet we are able to make good sense
of them. Once you take these sorts of examples into account, it becomes
clear that standard attempts to deal with the representation of space from
a single viewpoint merely scratch the surface of the problem of perceiving
pictorial space and point the way to further work to come.

We hope that the reader will not only enjoy the diverse discussion of pic-
torial space found in this collection, but that the volume’s chapters will be
an inspiration to continue the theoretical and empirical work advanced.
PART I
THE DUAL NATURE OF PICTURE PERCEPTION
This page intentionally left blank
Chapter 1
In Defense of Seeing-In

Richard Wollheim

When we look at a picture, by which I mean a representational picture, we


see something in it. That is my starting point. So when we look at Édouard
Manet’s Emilie Ambre (figure 1.1 and color plate 1), we see a woman in it.
In saying that we see a woman in it, I refer to a certain feature of our
perception of pictures, and this feature has important theoretical conse-
quences. Because of this feature what two well-known groups of theorists
find to say about pictorial perception is false. It is not the case that the rep-
resentational content of Manet’s painting, or the fact that it represents
Madame Brunet, is something that we infer from our experience of the
marked surface: for we also have an experience, a visual experience, of
her. Furthermore it is not the case that the visual experience we have of
Madame Brunet is such that, in the absence of information to the con-
trary, we would believe that we were seeing her face to face. When we look
at the painting, we see Madame Brunet, but the way we see her, or the mode
of perception, is such that we say that we see her in the picture. So, repre-
sentational content is experiential, but it is not the product of illusion.
Is there something that we can say directly about that feature of picture
perception, or seeing something in a picture, which carries with it this dou-
ble load of theoretical consequence?
I believe that we can, and, to pick out this phenomenological feature, I
have, for a number of years, used the term “twofoldness.” When I look at
the Manet, my perception is twofold in that I simultaneously am visually
aware of the marked surface and experience something in front of, or
behind, something else—in this case, a woman in a hat standing in front
of a clump of trees. These are two aspects of a single experience. They are
not two experiences: they are not two simultaneous experiences, as I used
4 Richard Wollheim

Figure 1.1
Edouard Manet, Emilie Ambre. Philadelphia Museum. See plate 1 for color
version.

to believe, nor are they two alternating experiences, as Ernst Gombrich


has claimed in Art and Illusion (1956).
It should not be hard to see how twofoldness refutes the two theories
of representational content that I have referred to. That, in looking at
Manet’s Madame Brunet, I experience a woman in a hat falsifies the infer-
ential view, and that I am visually aware of the marked surface refutes the
illusion view.
If we use the term “seeing-in” to identify the appropriate perceptual
mode to adopt toward representational pictures when we view them as
representational pictures—as opposed to some other way, such as regard-
ing them exclusively as paint samples or exclusively as props in some game
of memories—then my further claim is that twofoldness is an essential
5 In Defense of Seeing-In

feature of seeing-in. (In contrasting viewing pictures as representations


and viewing them some other way, I am leaving open the question
whether, once we have succeeded in seeing something that was intended to
be a representation as a representation, we are able to view it any other
way. I have serious doubts on this score.)

If we think of twofoldness as built into seeing-in, then we can use the


notion of seeing-in in two different ways.
In the first place, we can use the notion of seeing-in to provide a defini-
tion of a representational picture. A representational picture is a picture
that requires, or calls for, seeing-in. What frees this definition from trivi-
ality is that although representation can be explained through representa-
tion, we do not need representation in order to identify seeing-in. On the
contrary, seeing-in precedes representation both logically and historically.
We can see things in objects that were neither intended to be, nor are
believed by us to be, representations. Looking at a stained wall in Chicago
1948, we can see in the wall a boy carrying a mysterious box (figure 1.2).
And I am sure that our ancestors engaged in such diversions before they
thought of decorating the caves they lived in with images of the animals
they hunted.
Understood in this way, or through the perceptual skill of seeing-in, the
category of representational picture includes both figurative and abstract
pictures. It does not however include all figurative and abstract pictures,
for there are two comparatively restricted groups of pictures, one figura-
tive and one abstract, that do not, for what are complementary reasons,
call for twofoldness, and hence are not representational. On the figurative
side, there are trompe-l’oeil pictures (figure 1.3), for they do not call for
awareness of the marked surface; and, on the abstract side, there are those
pictures which aim at flatness (figure 1.4), for they do not call for us to
experience one thing in front of the other.
Second, we can use the notion of seeing-in to identify the representa-
tional content of a picture. A painting represents whatever can be cor-
rectly seen in it.
The point of “correctly” is to indicate that, whereas with nature, there
is no impropriety in anything that we may happen to see in, say, a stained
wall, and with a Rorschach test the only impropriety is to insist on seeing
nothing rather than something in the fabricated inkblots put in front of
us, there is, once representation arrives on the scene, one thing, just one
6 Richard Wollheim

Figure 1.2
Aaron Siskind, Chicago 1948.

thing that we do right to see in its surface. We do wrong—though the


wrong may be committed in play or jest, and not in error—if we see
anything else in its surface. When Marcel Proust went to the Louvre and
saw his friends from the Faubourg in the paintings of the old masters—
for instance, the Marquis du Lau in a remarkable double portrait by
Domenico Ghirlandaio (figure 1.5)—he was playing a game, and the game
was to see what he was not supposed to see. Once seeing-in is transposed
from nature to art, or at least to artifactuality, it gains a standard of
correctness.
I would make a further claim to the effect that this standard is set by the
fulfilled intentions of the artist. And the intentions of the artist are fulfilled
when, working under the causal influence of certain psychological factors,
which have come to be called in this particular theoretical context “inten-
tions,” he has so marked the surface of his painting that it is possible for
7 In Defense of Seeing-In

Figure 1.3
Alexandre I. Leroy de Barde, Reunion of foreign birds placed within different cases.
Nineteenth-century watercolor. Inv.: 23692. Photo: Michèle Bellot. Copyright
© Réunion des Musées Nationaux / Art Resource, NY. Louvre, Paris, France.
8 Richard Wollheim

Figure 1.4
Barnett Newman, Vir Heroicus Sublimis. 1950–51. Oil on canvas, 751183 55_1759 41 55
(242.2 _513.6 cm). The Museum of Modern Art, New York. Gift of Mr. and
Mrs. Ben Heller. Photograph © 2001 The Museum of Modern Art, New York.

a properly sensitive, properly informed, spectator to have an experience


that tallies with those factors. “Tally” here is a broad term, and its specific
application depends on the type of psychological factor, or intention, that
is seeking fulfillment. When the artist has a representational intent, which
is our current concern, this intention is fulfilled when it is possible for a
properly sensitive, properly informed, spectator to see in the surface that
the artist has marked the object that he intends to represent. Likewise the
artist’s representational intentions are unfulfilled when it is impossible for
an appropriate spectator to see that object in the picture. What makes
what an appropriate spectator can or cannot see in a picture relevant to
fulfillment of the artist’s representational intention is that experiences of
seeing-in “tally with” representational intentions.

Within the account that I have offered of how we identify a painting’s


representational content, there is a point that needs elaboration. I have
equated a painting’s representational content with what can be correctly
seen in it, and I have equated what can be correctly seen in a painting with
an experience that conforms with the artist’s intentions. In other words,
the artist’s control over the content of the picture that he is making is
9 In Defense of Seeing-In

Figure 1.5
Domenico Ghirlandaio (1448–94). Portrait of an old man and his grandson. Copy-
right © Alinari / Art Resource, NY. Louvre, Paris, France.

limited in this respect: that it is out of the things that can be seen in his
picture that his intention determines what the picture represents. If some-
thing cannot be seen in a painting, it forfeits all chance of being part of its
content.
So the question arises, How can we determine, for a given picture,
whether something can be seen in that picture? What are the limits of
visibility in a picture?
The question is clearly ambiguous between an extensional and a nonex-
tensional reading. Since there is little doubt but that when we ask whether
10 Richard Wollheim

something can be seen in a picture, the “something” is clearly to be under-


stood in a nonextensional way, we should reformulate the question thus:
How do we determine, for a given picture, the range of things that some-
thing that can be seen in that picture can be seen as?
One proposal for answering that question is this: In a given picture, we
pick out a certain object that can be seen in it. Let us imagine that we
merely pick it out as “that thing.” Then people in turn propose vari-
ous concepts that apply to the thing in question, and what we have to
decide for each such concept is whether it affects how we see the object or
whether it merely affects what we say about it. If for a given concept,
and the concept could be one that had independently occurred to us, the
answer is positive, then we can see the object as falling under that concept.
If the concept is f, then we can see the object as f, and fs are visible in that
picture.
In an article of recent years, I applied this test for visibility in a picture
in the following way: We look at a painting showing a figure seated
amongst classical ruins (figure 1.6), and someone starts to ask questions of
us, to which we must answer, Yes, or No. So, for example, Can you see
those columns as having been thrown down? Yes. Can you see those
columns as having been thrown down hundreds of years ago? Yes. Can
you see those columns as having been thrown down hundreds of years ago
by barbarians? (with some difficulty) Yes. Can you see them as having
been thrown down hundreds of years ago by barbarians wearing wild
asses’ skins? (with little difficulty) No. Though we are perfectly ready to
believe that the barbarians whose handiwork our eyes are prepared to
acknowledge in this picture did indeed wear wild asses’ skins, there is no
way in which this belief, or the concept that figures in it, can help us to
structure our perception of the painting, or can affect how we see the
columns. So, of the range of concepts put to us in this test, it is the only
one that is not instantiated by something visible in the picture.
Those who are impressed by the modularity of the faculties are likely to
resist this test and what it is supposed to show, but to them all I can say is
that, whatever credence we might give to the role of modularity in per-
ception in general, there is obviously a level of complexity above which it
doesn’t apply, and there is reason to think that picture perception lies out-
side its scope. The reason can be brought out by our considering pictures
of women. Now, of some of these pictures, such as Manet’s Emilie Ambre,
when we further ask, What woman is it of? there is an answer to this ques-
11 In Defense of Seeing-In

Figure 1.6
Nicolas Poussin, French, 1594–1665, Landscape with Saint John on Patmos, 1640,
oil on canvas, 100.3 _136.4 cm, A. A. Munger Collection, 1930.500 © The Art
Institute of Chicago. All rights reserved.

tion, whereas with the rest, such as Manet’s La Prune (figure 1.7), there is
no answer to this question. To generalize, some pictures are of particular
things or events, and others are of things or events merely of some partic-
ular kind.
The explanation for why we classify the two paintings in different ways
is not that there is in reality such a person as Madame Brunet and that
Manet painted his painting by painting her, whereas there is no particular
person whom La Prune represents, though all this is in fact true. For, in
the first place, I believe (contra Nelson Goodman) that we should put
Jean-Auguste-Dominique Ingres’s Jupiter and Thetis in the same category
as Madame Brunet even though Jupiter and Thetis are not real persons.
Second, to invoke this kind of explanation at this stage would be to go
12 Richard Wollheim

Figure 1.7
Edouard Manet, La Prune (Plum Brandy), c. 1877, oil on canvas, .736 _ .502
(29 _19 43 ); framed: .876 _.641_.057 (34 12_25 14_2 14). Collection of Mr. and
Mrs. Paul Mellon. Photograph © 2001 Board of Trustees, National Gallery of
Art, Washington.
13 In Defense of Seeing-In

into reverse about how a representational picture comes by its content.


Accordingly, if we really do think that there is a division within paintings
of the kind I propose, according to which Emilie Ambre and Jupiter and
Thetis lie on one side of the line and La Prune lies on the other, this must
be because there is a difference, a perceptual difference, between an expe-
rience of a painting that we would naturally report by saying “I see a par-
ticular woman in that picture” and an experience of a painting that we
would more naturally report by saying “I see merely a woman in that pic-
ture.” And, if this is so, then one experience must be laden with referential
thought and the other laden with nonreferential thought, and this would
in turn show that, in the domain of picture perception, perception exem-
plifies a rather remarkable degree of permeability by thought.

I return to the claim that when I look at a picture appropriately, the


twofoldness that qualifies my perception is not a matter of my having two
experiences, each one of a familiar kind, which can also be had outside the
aegis of seeing-in. Rather I have one experience with two aspects, and this
is distinctive of, if not unique to, seeing-in.
I have said that some philosophers deny or have denied this idea out-
right. But other philosophers, and also (or so it seems to me) some per-
ceptual psychologists with interests in this area, in effect assume its falsity.
Whether this is so or not, let me set up a straw man who does assume two
separate experiences, and I want to consider how this assumption can
grossly distort something that he might be inclined to think of as “the
problem of picture perception.”
If someone holds that looking at a picture is not characterized by
twofoldness but is constituted of two paired experiences, each of a kind
familiar to us from outside the perception of pictures, one of which pre-
sents the viewer with something two-dimensional, the other with some-
thing three-dimensional, then the person is likely to think that the problem
of picture perception is how to explain the occurrence of a problematic
experience, that is, the experience of something three-dimensional when
confronted solely by a two-dimensional surface. That being so, the natu-
ral next step is to look for the explanation of this experience within the
experience paired with it: namely, the experience of looking at the marked
surface, and seeing it as such.
14 Richard Wollheim

Our straw man goes wrong in locating the explicandum of his inquiry
where he does. For, when we look at, say, Emilie Ambre, it is not the case
that we have an experience of the sort that would, in the absence of ancil-
lary evidence to the contrary, lead us to believe that we were seeing Emilie
Ambre face to face. But might our straw man not yet be right in what he
takes to be the explicans? He might be. It might be the case that, whenever
we saw something in a marked surface, we were led to do so by a prior
experience of the surface for the duration of which we inhibited ourselves
from seeing something in it. But I am skeptical that this is so: I cannot
readily believe that every occasion on which we see something in a marked
surface is preceded by our seeing that surface without seeing anything in
it. So the latter kind of experience cannot furnish an explanation of the
former kind.
A different kind of explanation of why we see something in a marked
surface, though the two can get confused, is an appeal not to some way in
which we see the surface but to some way in which the surface is. The way
the surface is, in such cases, is not necessarily something of which we are
aware, but, unless it were as it is, we would not see in it what we do.
It is sometimes argued that, in any account of what we see in a surface,
there is no room for any feature of the surface to play a role unless this
feature is itself visible. But this betrays a confusion between two kinds of
account. For it is certainly the case that in an analysis of what it is to see
something in a marked surface, a feature of the surface of which there is
no awareness is out of place. This is because an experience can only be
analyzed in terms of constituent experiences. But when it comes to an
explanation of what it is to see something in a surface, the situation is
different. For we can explain an experience in terms of other experiences,
but also in terms of features of the environment to which the experience is
sensitive, even though the person might have no awareness of them as
such. In the second case, the explanation does not double as an analysis.

If we do think that it is possible in principle to explain how we see a


woman’s head in a painting by citing objective features of the painted
surface, it is important to recognize that because of the vast diversity of
ways in which a woman’s face can be represented or the vast diversity of
15 In Defense of Seeing-In

painted surfaces in which a woman’s face can be seen, the explicandum


in such cases will be not something of the generality of the representation
of a woman’s face but something of the particularity of Raphael’s depic-
tion of a woman’s face, or Guercino’s depiction of a woman’s face, or
Picasso’s depiction of a woman’s face, or Willem de Kooning’s depiction
of a woman’s face. Some might feel that this would leave the central ques-
tion of representation, why these are all representations of the same thing,
unanswered. But anyone sensitive to what a painter is doing when he
paints a woman’s face might prefer not to raise this question.
This page intentionally left blank
Chapter 2
Conjoint Representations and the Mental Capacity for Multiple
Simultaneous Perspectives

Rainer Mausfeld

Mens videt, mens audit: Cetera surda et coeca.


It is the mind that sees and the mind that hears;
the rest are deaf and blind.
—Epicharmos1

Commonsense taxonomies were, inevitably, the origin from which the


natural sciences, at their earliest stages of development, derived their
categorizations of phenomena. This can be witnessed by the classical
division of physics into, for example, optics, acoustics, theory of heat, and
mechanics. During the process of its theoretical development, physics
became increasingly divorced from these kinds of classifications and rather
grouped phenomena in accordance with its own internal theoretical struc-
ture (the classical theory of heat, for instance, disintegrated into statistical
mechanics, on the one hand, and electrodynamics, on the other hand). In
perceptual psychology corresponding pretheoretical classifications of phe-
nomena are mirrored in the standard textbook organization in terms of
salient perceptual attributes, such as color, depth, size, or form. In this
field, it will likely prove to be more difficult than it was in physics to
dispense with commonsense classifications of phenomena and to instead
follow lines of theorizing that are traced out by the development of suc-
cessful explanatory accounts. This difficulty is due to the power that the
phenomenal appearance exercises over our way of theoretically grouping
perceptual phenomena. Although we are well aware that commonsense
taxonomies are an inapt guide for the endeavor to achieve, within the
framework of the natural sciences, a theoretical understanding of the
mind, we are held captive by the appearances. We are inclined to believe
that perception works in the way it phenomenally appears to us and that
a theoretically fruitful classification of phenomena basically follows our
18 Rainer Mausfeld

commonsense psychological intuitions. It was such pretheoretical com-


monsense taxonomies that in perceptual psychology brought forth sub-
fields such as picture perception or color perception.
For someone, like me, who is working in the field of color perception,
picture perception does not appear as a natural field of inquiry, all the
more so because hardly any other two fields in visual perception are as
remote in their approaches and their theoretical frameworks as picture
perception and color perception. Nevertheless, I believe that phenomena
in these fields share, on a more abstract level, structural similarities that
appear to point to deeper principles of our mental architecture. This is
hardly surprising, because there is no reason to expect that classes of per-
ceptual phenomena that arise in the context of certain artifacts, such as
pictures, and classes of phenomena that pertain to certain phenomenal
attributes, such as color, will survive as theoretically useful categories
when more successful explanatory accounts of perception have eventually
been developed. Rather, such accounts would result in a classification of
phenomena that is determined by the actual internal principles underlying
perception, whatever these may turn out to be.
The most promising general approach to perception appears to me to
be one that follows ethological lines of thinking. Such approaches have,
notably when couched in computational terms, already yielded intriguing
explanatory frameworks of promising range and depth. From an etholog-
ical perspective, the core task of perceptual psychology is to investigate
the structure of perceptual representations and the nature of the repre-
sentational primitives on which it rests. With respect to color perception,
such an approach has brought forth what I believe to be interesting
theoretical speculations about the nature of the representational primi-
tives and about the internal structure they give rise to. The internal
principles that are suggested by these speculations and that we are only
beginning to understand are fairly general ones that crosscut common-
sense categorizations of phenomena. Although not much is presently
known about the structure of the representational primitives to which
these principles refer, some of their general properties are suggested by
interesting structural similarities that certain phenomena from various
domains in cognitive science appear to share. Among these phenomena is
our ability to perceptually deal with the so-called dual character of pic-
tures. This ability is, I believe, part of a general structural property of our
basic cognitive architecture that refers to the handling of conjoint repre-
sentations over the same input. In various domains and at different levels
19 Conjoint Representations

of the cognitive system, we can encounter phenomena that are likely due
to the internal handling of multiple conjoint and often competing repre-
sentations. From this point of view, picture perception and its dual char-
acter are a special instance where we exploit these given capacities in the
context of human artifacts. Before arguing for this view in greater detail,
I will briefly delimit the topic of my inquiry.
In picture perception, more than, say, in research on stereo vision or
color coding, the tension looms large between what are to be considered
universal and what cultural and conventional aspects, a tension that mir-
rors and maintains the time-honored distinctions that placed physis and
ethos, nature and convention, essential and accidental properties in well-
nigh irreconcilable opposition to one another. The corresponding issues
are a matter of much debate between—to use some fashionable jargon—
universalists and cultural relativists. Outside some areas of vision and
language, little of substance is presently known about where to draw a
dividing line between universal aspects and aspects of cultural variation,
individual plasticity or learning history. However, the entire idea of cog-
nitive science rests on the assumption that some such distinction can be
drawn at all. This also holds for inquiries centering on picture perception.
The perception of pictures, on the basis of multiple pictorial components
of very different status, involves highly complex interactions of our per-
ceptual faculty and various interpretative faculties, which are presently
not well understood. These interactions give rise to a high degree of cul-
tural variation. I shall deliberately ignore the cultural dimension of picture
perception and, with respect to the so-called dual character of pictures,
focus on structural elements of perception that seem to be part of our
basic cognitive endowment.

2.1 THE DUAL CHARACTER OF PICTURES

Pictures and pictorial representations, though highly impoverished two-


dimensional abstractions of what is depicted, can evoke strong perceptual
impressions of objects, spatial relations, or events within us. A dominant
theme in the field of picture perception has been the issues centering
on notions of perceptual space and the extent to which corresponding
percepts can be evoked by features of pictorial representations, notably
through linear perspective (cf. Haber 1980; Rogers 1995). During the
Renaissance an increasing interest emerged in techniques of linear per-
spective. This was motivated by the artists’ desires to imitate nature and
20 Rainer Mausfeld

to achieve “visual truth” in their paintings. This idea gave rise to related
inquiries into artistic techniques for the evocation of space and, in partic-
ular, into techniques for creating geometrically correct two-dimensional
pictorial representations on a canvas of the three-dimensional layout of
the pictured scene. In such investigations, as Martin Kemp observed, “the
eye figures little, the mind features even less” (1990, p. 165). Rather, what
was to be accomplished was “the demonstration of an internally consis-
tent system of the spatial elements in a picture and, above all, a proof that
the system rested upon non-arbitrary foundations” (ibid., p. 11). The can-
vas was regarded as a window, often referred to today as an “Alberti win-
dow,” through which the painter views the world and that intersects the
painter’s visual cone (Lindberg 1976). This conception gave rise to the
idea that a realistic appearance of depth and space can be achieved in pic-
tures by mimicking the exact geometrical relations in the structure of light
that reaches the eye from a three-dimensional scene. Consequently, a sys-
tem of construction rules, in the sense of artistic-engineering techniques
for the purpose of creating, on a flat canvas, pictorial representations that
induce a strong appearance of depth in the observer, gained prominence
in Renaissance art.2 Although these artistic techniques later joined with
ideas on geometrical processes of image formation in the eye, their use and
development were primarily shaped by considerations internal to the com-
plex variety of cultural purposes underlying artistic productions. However,
for the endeavor to imitate nature and to achieve visual truth in two-
dimensional representations of the world, the importance of rules for lin-
ear perspective is on a par with those for simulating the effect of lights
and the interaction of light and objects by using spatial pigment patterns
on a flat surface (Schöne 1954). It is a historically contingent development
of art history that resulted in linear perspective, rather than other aspects,
first gaining prominence in the context of artistic attempts to imitate
nature.
The notion of a dual character of pictures basically refers to the phe-
nomenon that pictures can generate an in-depth spatial impression of the
scene depicted while at the same time appearing as flat two-dimensional
surfaces hanging on a wall. Albert Michotte (1948/1991) recognized the
challenge that this kind of phenomenon poses for perception theory.3 A
description in terms of a perceptual conflict between the perceived flat-
ness of the picture’s surface and the perceived depth of what is depicted
captures only a small fraction of the perceptual enigmas involved. By
careful phenomenological observation when viewing pictures, one can
21 Conjoint Representations

easily gather indications of how difficult it is to precisely describe what


the percepts are and thus what the perceptual achievements are in these
situations.4 I will deliberately leave out many of the problems that we
encounter in describing exactly the percepts or the perceptual achieve-
ments in corresponding situations of viewing pictures. I will, rather, dis-
tinguish only two general problems to which the notion dual character of
picture seems to refer, which I will address in turn. The first is the prob-
lem of cue integration with respect to depth or, more generally, spatial
representations.5
The second is a problem that seems to me to be even more complex and
much deeper, namely, the problem of what I refer to as conjoint represen-
tations over the same inputs and our ability to smoothly handle them. By
conjoint representations I will, in an intuitive and tentative way, denote
two or more representational primitives of the same type or of differ-
ent types that exploit the same input properties in different and interde-
pendent ways. The special case of competing conjoint representations is
furthermore characterized by the property that the parameters of one
representational primitive that relate to a certain aspect of the input are
antagonistically coupled with parameters of the other representational
primitive that refer to the same aspect. As an illustration think of a sur-
face viewed under some chromatic illumination: the incoming color signal
is internally exploited in terms of two components, that is, “surface color”
and “illumination color” (to be understood as internal and not as physi-
cal concepts); the extent to which a “surface” representation exploits the
incoming color signals in terms of its color parameter antagonistically
constrains the values that an “ambient illumination” representation can
assign to its color parameter, and vice versa.
The two issues of the integration of conflicting cues with respect to a
specific representational primitive, on the one hand, and a dual represen-
tation that results from conjoint representations over the same inputs, on
the other hand, are often conflated.

2.2 CONFLICTS IN CUE INTEGRATION WITH RESPECT TO DEPTH OR


SPATIAL REPRESENTATIONS

The lines along which we can theoretically explore the dual character of
pictures considered as a case of cue integration are comparatively well
explored in visual psychophysics. Although problems of cue integration
are intricate enough we have both a wealth of experimental results and
22 Rainer Mausfeld

subtle mathematical-modeling tools, such as the Bayesian framework,


which has proved a fruitful basis for dealing with these issues. All this is
well known, and I will only briefly review some of the relevant evidence
in order to better differentiate the second problem of conjoint representa-
tions from issues of cue integration.
We know from psychophysics that the depth one experiences, that is,
the apparent variation in surface relief and the relative location of objects
in 3-D space, is constructed from multiple sources of evidence. These cues
can carry different weights with respect to internal spatial representations,
weights that can be but are not necessarily in line with natural ecological
constraints. For instance, occlusion carries a strong internal weight but
provides only weak constraints for spatial representations, namely, that
the depth on one side of the border is greater than on the other side.
Another example is shading, which provides information about the sur-
face normal at each location. Even cues that are unrelated to depth per se
can be used to disambiguate other cues, such as spatial frequency content
for 3-D-shape-from-texture cues. If different sources of evidence favor
different and mutually incoherent spatial representations of the same
input, the visual system often has some kind of default preference for solv-
ing this cue conflict. It does this without providing any phenomenal access
to alternative representations. A case in point is the Ames room, where
systematic manipulations in linear and texture perspective cues result in
estimates of depth and size based on a cue integration that vetoes or
ignores cues of familiar size; the cue integration seems rather to be in line
with internal heuristics such as “lines that are nearly parallel in the image
are parallel in 3-D,” or “lines that meet at a common vertex in the image
also meet at a common vertex in 3-D.”
In picture perception we find a plenitude of ways to evoke space and
depth on a two-dimensional surface that mirror the plenitude of depth
cues. We encounter a similar situation for the integration of conflicting
cues, namely the integration of stereo disparity information and various
monocular depth cues.6 Although the depth information of the picture
surface is perfectly in line with the depth information from the frame or
the surrounding wall, the scene depicted can invoke the impression of
being phenomenally extending in depth.
In principle we could conceive a visual system that integrates these cues
in such a way that the stereo information completely dominates and, in
the case of incoherence, vetoes monocular information, which phenome-
nally would result in a reduction of the—already not very strong—3-D
23 Conjoint Representations

vividness of the scene depicted to zero, that is, a flat picture without
any indication of surface orientation or relative 3-D locations of objects.
Interestingly enough, our visual system uses almost the opposite strategy:
it is a well-known result from visual psychophysics that monocular cues
often dominate the resulting 3-D interpretation over stereopsis, even at
close range where stereopsis is most accurate. I will mention only a
few corresponding studies. For instance, in a pioneering study, Walter
Schriever (1925) found that perspective alone as well as occlusion could
overrule disparity information. Schriever also made a wealth of care-
ful observations about attentional effects and vagueness and instability,
as well as individual differences. Mümtaz Turhan (1937) found that in
center-surround situations, brightness gradients of opposite direction can
result in perceptual impressions that violate the physical depth relations
of infield and surround as provided by disparity information and motion
parallax (infields whose brightness gradient has the same direction as that
of the surround appear to lie in the same depth level as the surround,
whereas they appear to lie in a different depth plane and often look bent
in the case of opposite brightness gradients). Turhan observed that phys-
ically incompatible depth interpretations of the infield can occur at the
same time and often are accompanied by some kind of vagueness of the
perceptual impression.
John Yellott (1981) showed that an inside-out face—a mold of a face—
looks right side out as long as shading is present, despite the presence of
contradictory disparity information as provided by a random-dot stereo-
gram projected onto the surface of the mask. If the mold is presented
solely as a random-dot stereogram with no shading, it is seen inside
out, that is, consonant with the disparity information. Another interest-
ing example is provided by Kvetoslav Prazdny (1986), who described a
random-dot stereo cinematogram that portrays a flat object in front of
a background that changes its two-dimensional shape consistent with a
three-dimensional rotating wire object, while at the same time the binocu-
lar disparities are incompatible with the relative depth information speci-
fied by the image motions. Due to the appearance of three-dimensionality
in these displays he concluded that the kinetic depth effect effectively
vetoes the stereo disparity cue with respect to the shape of the object
(however, disparity determined the position of the object with respect to
the background).
A particularly effective demonstration of the monocular influence over
stereo information was provided by Kent Stevens and Allen Brookes
24 Rainer Mausfeld

Figure 2.1
Stimulus configuration used by Stevens and Brookes (1988). The lines are stereo-
scopically presented as being coplanar, that is, they increase linearly in disparity
from left to right. The 3-D impression, however, is of a corridor extending in
depth, as suggested monocularly.

(1988; see also Stevens, Lees, and Brookes 1991). The lines in the stereo-
scopically presented figure 2.1 are coplanar, that is, they increase linearly
in disparity from left to right. The 3-D impression, however, is of a corri-
dor extending in depth, bordered on either side by columns of vertical
lines or stakes. In the stereo-apparatus the innermost lines on either side
of the vertical meridian had stereo disparities of <11 minutes of arc; the
outermost lines had disparities of <515. It is remarkable that the line
with :115 disparity appeared more distant than the line of disparity +515.
Stevens and Brookes found empirical evidence that stereopsis extracts
3-D surface information only where the second spatial derivatives of dis-
parity are nonzero, corresponding to loci where the surface is curved,
creased, or discontinuous.
We now can directly apply their result to the situation of picture per-
ception. A picture hanging on a wall has no local disparity differences over
the picture surface but only a continuous uniform gradient of disparity
25 Conjoint Representations

indicating a flat although extended surface. Accordingly, because of the


specific properties of the human visual system for integrating binocular
and monocular sources of depth evidence, as exemplified by Stevens
and Brookes’s results, stereopsis is particularly weak under these condi-
tions. Our phenomenal spatial impression is therefore—in the presence of
monocular depth cues such as occlusion, texture, shading, or perspective—
determined by these monocular cues. Thus, in picture perception, we take
advantage of the internal coding property that depth is derived from
disparity only where the surface exhibits continuous curvature or sharp
discontinuities, because some binocular disparity information must be dis-
counted in picture perception to interpret drawings at all.
I will not dwell further on the cue integration aspect of the dual char-
acter of pictures (for a more recent study of cue integration with respect
to depth, see Landy et al. 1995). What I wanted to show is that the conflict
between stereo information and monocular depth cues, as well as the cor-
responding observation that the flatness of the picture plane does not
impede the depth impression of the scene depicted, is simply a special case
of the specific way our visual system integrates disparity information with
other depth cues, as mirrored in a great variety of psychophysical results.

2.3 CONJOINT REPRESENTATIONS IN PICTURE PERCEPTION

I will now turn to the second aspect of the notion of dual character of pic-
tures, namely, the issue of conjoint representations over the same input
and our ability to handle them smoothly. This issue is much more puzzling
and much less understood than the conflicting cue aspect. I shall first
describe the corresponding kind of phenomenon, as referred to in the lit-
erature on picture perception. Like the cue integration aspect, it is, in my
view, not specific to picture perception but rather a general and essential
property of our cognitive organization.
In picture perception, we can simultaneously have the phenomenal
impression of two different types of objects, each of which seems to thrive
in its own autonomous spatial framework,7 namely, on the one hand, the
picture surface as an object—with corresponding object properties such as
orientation or depth—and, on the other hand, the depicted objects them-
selves with their idiosyncratic spatial properties and relations. We seem to
have two mutually incompatible spatial representations at the same time;
at least in the sense that they are available internally and we can, without
any effort, switch back and forth between them.
26 Rainer Mausfeld

Before venturing some more speculative ideas about some general prop-
erties of our cognitive architecture on which this ability rests, I will list
some observations concerning this dual nature of pictures that I consider
to be of particular relevance.
In an aside, in order to avoid potential misunderstandings I would like
to emphasize that pictorial art in general cannot simply be understood
as a kind of frozen optical array or a static boundary case of the optical
structure of the input from a scene. Pictorial art is much richer than nat-
uralistic artistic productions and serves a great many different symbolic
functions. Pictures are not surrogates for scenes, nor can they be sub-
jected to a criterion of some absolute notion of veridicality (whatever
that means) with respect to the scene depicted. Thus, those aspects of pic-
ture perception that we potentially can understand from core perceptual
principles—and, as I said in the beginning, it is only this part that I will
address here—are, from the perspective of cultural studies, the least inter-
esting aspect of picture perception. What we usually refer to when we
talk about picture perception are symbolic interpretations at various
levels and thus aspects that pertain to highly complex interactions of our
perceptual faculty and various interpretative faculties. Within the frame-
work of the cognitive sciences we virtually know next to nothing about the
cognitive principles underlying these achievements. Let us look at four
types of observations that are of particular theoretical relevance:
1. A continuous path of transitions exists from a view of a real 3-D scene
to a scene as depicted (or abstracted) on a canvas.
We can easily construct a continuous path from the view of a real 3-
D scene to a real 3-D scene viewed as a kind of frozen Alberti window
to a photo of this frozen optical array and to a highly reduced drawing
of the relevant contours of the scene. This allows us to experimentally
investigate all sorts of transitions and boundary conditions in picture
perception.8
2. We can phenomenally accentuate one or the other aspect and switch
back and forth in an effortless way.
This aspect is phenomenally so conspicuous and striking that we usual-
ly do not pay much attention to it. It is at the core of what we mean when
we refer to the dual nature of pictures. Though such switches are corre-
lated with depth aspects, they actually pertain to the entire perceptual
organization of the visual field and thus to attributes such as shape, or
shading and brightness gradients. A wealth of observations pertaining to
this kind of phenomenon have been reported in the literature, a wealth
27 Conjoint Representations

that contrasts oddly with the silence about what to make of these obser-
vations theoretically.
Ernst Gombrich (1982) made the important observation that one has
to achieve the proper mental attitude to take full advantage of the capac-
ity to switch back and forth between the reality of the picture as an object
and the reality of the depicted objects. Because of this, people at earlier
stages of cultural development regularly seem to have problems seeing
what is depicted in a photo. For instance, Jan Deregowski, Muldrow, and
Muldrow (1972) reported that people from a remote Ethiopian tribe when
presented with a drawing of an animal would pay attention to the charac-
teristics of the drawing paper but would ignore the picture; they exhibited
a complete inattention to the content of the representation while concen-
trating on the medium. Several others have reported essentially the same
observation. When people at earlier stages of cultural development regu-
larly have problems in seeing what is depicted on a photo, they have
not yet attained the ability to exercise what Gombrich called the proper
mental attitude and thus cannot fully exploit a given cognitive capacity in
the case of previously unknown artifacts. However, as Margaret Hagen
and Rebecca Jones concluded from a review of corresponding studies,
“this coexistence of information poses few problems even for the naive
observers when pictures represented only single solid objects. There is no
evidence whatsoever that any group of people see pictures of faces, cups,
hunters, antelopes or elephants as flat ‘slices of life,’ as it were” (1978, p.
192). Many other interesting regularities were found with respect to the
ability to simultaneously handle both types of reality, as it were. For
instance, outline drawings present fewer difficulties than photos to naive
observers; thus contour information seems to be of greater importance
than texture.9
Corresponding observations are, of course, not confined to picture
perception, but pervade psychophysics and perceptual psychology. For
example, in the study mentioned previously, Stevens and Brookes (1988,
p. 383) made the observation that experienced stereo observers can also
discern the true stereo depth of the component lines with scrutiny, as if
they can selectively disregard the monocular depth interpretation.
3. The “realities” of pictures as objects and depicted objects bear dif-
ferent amounts of internal computational relevance and phenomenological
vividness.
We can switch back and forth between these two kinds of spatial repre-
sentations, but this is not a switch between a 2-D representation (within
28 Rainer Mausfeld

some 3-D representation) and a different full-fledged 3-D representation.


Rather, the spatial representation of the scene depicted is phenomenally
quite shallow. In a sense we could regard it as a phenomenal analog to
David Marr’s 2 12 D spatial representations, that is, what we perceptually
experience are local surface orientation, distance from viewer, or discon-
tinuities in depth and discontinuities in surface orientation, without either
the phenomenological vividness or the other internal properties of a full-
blown 3-D representation. There are, of course, several other aspects
that are responsible for the reduced 3-D vividness of the scene depicted,
notably a lack of cues provided by motion, and the ranges of color and
luminance contrast, which are much narrower than those of real scenes.
Even if we attend to the two-dimensional picture surface we may expe-
rience the depicted object, say, a line drawing of a cube, in a mandatory
way as three-dimensional, and yet this 3-D spatial representation lacks
other crucial elements of a full-blown 3-D representation; for instance, it
would hardly fool us into trying to grasp and rotate the cube.10 Both types
of representation exhibit therefore quite different kinds of anchoring with-
in the internal computational structure, and we can safely conjecture that
we have specific mechanisms subserving “flat” representations by ignor-
ing certain aspects of 3-D structure. Pieter Saenredam’s Interior of the
Saint Jacob Church in Utrecht (1642; figure 2.2 and plate 2) may serve as
another illustration. In this case the conflict with size relations in the spa-
tial representation of the picture viewed within the environmental context
already suffices to divorce the—otherwise mandatory—3-D interpretation
of the scene depicted from other internal-coding properties of full-fledged
3-D representations.
4. The two representations are not independent but interlocked.
When we look at a picture like the one displayed in figure 2.2, it seems
that we can achieve a kind of autonomous spatial representation of the
scene depicted that is detached from the normal spatial representation of
our environment, including the picture surface plane. In other words, we
seem to use our faculty for spatial representations to emulate certain of
its achievements in a restricted local framework.11 As a result, the internal
output of such an emulation is a partly autonomous, though compar-
atively faint, local spatial representation within the canonical spatial
representation of our environment. In fact, both representations are nev-
ertheless internally interlocked; just how they are interlocked is poorly
understood. I will mention only two observations in this regard: first, in
many cases, we can experience that—at least locally—the spatial vividness
29 Conjoint Representations

Figure 2.2
Pieter Jansz Saenredam, Interior of the Saint Jacob Church in Utrecht, 1642.
Bayerische Staatsgemäldesammlungen, Alte Pinakothek, Munich. See plate 2 for
color version.
30 Rainer Mausfeld

gained by one aspect is lost by the other. Second, as Hagen and Jones
(1978, p. 194) also pointed out, “the space behind the picture plane is not
completely separated from the space of ordinary environment which sur-
rounds the picture.” For instance, Deregowski, Muldrow, and Muldrow
(1972) observed an interesting effect of a horizontal versus a vertical pres-
entation of a picture of a profiled standing buck. When the picture was
presented lying flat on the ground, most observers reported that the buck
was lying down; when it was presented vertically, they reported that the
buck was standing up.
The proposals and corresponding observations by Maurice Pirenne
(1970), James Farber and Richard Rosinski (1978), Michael Kubovy
(1986), and others that the surface characteristics of the picture have to be
available internally, in order to allow some kind of compensation process
that corrects distortions caused by an inappropriate viewing geometry,
also indicate that both representations are interlocked in complex and
poorly understood ways. In other areas, such as color or brightness per-
ception, we have a better theoretical understanding about how conjoint
representations are interlocked.
The structural perceptual properties that we can identify in theoretical
analyses of phenomena centering around the dual character of pictures
cannot simply be regarded as kinds of “perceptual irregularities” that are
due to encountering an artifactual situation. Rather, these phenomena
seem to point in a particularly conspicuous way to a general perceptual
capacity to deal with conjoint representations.

2.4 TRIGGERING AND PARAMETER SETTING: THE DUAL FUNCTION


OF SENSORY INPUTS WITH RESPECT TO REPRESENTATIONAL
PRIMITIVES

In perception theory we can roughly distinguish three aspects of archi-


tecture with respect to the relation of sensory inputs and internal rep-
resentations: (1) Several kinds of input properties are exploited by the
same kind of internal representation (e.g., computational theories of cue
integration), (2) the same input property is independently exploited by
several representations (e.g., in bees, color vision proper and wavelength-
dependent behavior coexist and subserve independent representations,12
cf. Goldsmith 1990), and (3) the same type of input property is exploited
by conjoint representations over the same input (i.e., the same input can
give rise to several different but interlocked output codes and to multiple
31 Conjoint Representations

simultaneous layers of representation). This last type of architecture can


be expected to play a prominent role in highly versatile and complex per-
ceptual systems that have to simultaneously subserve a great variety of
tasks. Such systems must internally have the outputs of many subsystems
available for purposes of a great variety of higher-order representations
and thus have to provide computational means to handle conjoint repre-
sentations over the same input.
In this section, I will briefly sketch a theoretical perspective on the
structure of perceptual representations that I believe to be promising, both
theoretically and empirically, for attempts to deal with conjoint represen-
tations. This approach is inspired by two general perspectives, which are
related in some of their core ideas. The first perspective refers to the ideas
underlying classic ethological approaches, notably of Jakob von Uexküll,
Konrad Lorenz, and Nikolaas Tinbergen, and by the extension of these
ideas to richer and more complex functions (e.g., Wehner 1987; Marler
1999; Gallistel 1998) than those studied in earlier years under the head-
ing of “innate release mechanism.”13 An ethological perspective has, in its
basic theoretical assumptions, also gained support from computational
approaches. The second perspective is Noam Chomsky’s (e.g., 2000) inter-
nalist inquiry into the nature of language and mind (which adhere to
the maxim that in rational inquiries into mental phenomena there is no
reason to deviate from the methodological principles routinely employed
in other domains of the natural sciences with respect to other “natural
objects”; a maxim that should be uncontroversial, yet, in the cognitive
sciences, assumptions to the contrary still remain highly influential).
Needless to say, the theoretical picture of the basic principles underlying
perception that has been emerging in corresponding studies is still very
skeletal and, of necessity, has to be based on considerable theoretical
speculation. Yet, an ethology-inspired internalist approach, which focuses
attention on the structure of the representational primitives underlying
perception, seems to provide a promising framework for asking novel
and potentially fruitful questions about the internal architecture of per-
ception. Such an approach is also less susceptible to the physicalistic trap
(Mausfeld 2002), which, in the history of perceptual psychology, has often
hindered the asking of appropriate questions.
In perceptual psychology, a wealth of empirical and theoretical evi-
dence has been marshalled by Gestalt psychology, Michotte’s “experi-
mental phenomenology”,14 ethology, studies of the newborns and young
children, and computational approaches, evidence indicating that the
32 Rainer Mausfeld

structure of internal coding is built up in terms of a rich set of representa-


tional primitives. According to the theoretical picture that has emerged
from corresponding studies, perception cannot be understood as the
“recovery” of physical world structure from sensory structure by input-
based computational processes. Rather, the sensory input serves as a kind
of sign for biologically relevant aspects of the external world that elicits
internal representations on the basis of given representational primitives.
(Thus, even “highly impoverished” sensory inputs can trigger perceptual
representations whose “complexity” far exceeds that of the triggering
stimulus and whose relation to the sensory input can be contingent from
the point of physics or geometry.) Although the sensory input is a caus-
ally necessary requirement for perceptual representations, the perceptual
computations triggered are under the control of an internal program
based on a set of representational primitives. They are representation
driven rather than stimulus driven.
With respect to human perceptual capacities, this theoretical perspec-
tive and the evidence on which it is based suggest distinguishing, as an
idealization, a sensory system from a perceptual system.15 Whereas the
sensory system deals with the transduction of physical energy into neural
codes and their subsequent transformations into codes that are “readable”
by and fulfill the needs of the perceptual system,16 the perceptual system
contains, as part of our biological endowment, the exceedingly rich per-
ceptual vocabulary in terms of which we perceive the “external world”
(whose relevant aspects pertain not only to physical and biological aspects
but also to mental states of others). Furthermore, the perceptual system
provides the computational means to make these perceptual concepts
accessible to higher-order cognitive systems, where meanings are assigned
in terms of external world properties. The sensory system interfaces
directly with the motor system (this interface is evolutionarily an old one)
as well as with the perceptual system, whereas the perceptual system inter-
faces with the motor system and higher-order cognitive systems. Because
the sensory system provides, in terms of its physicogeometrical vocabu-
lary, the cues that the perceptual system exploits in terms of its concep-
tual structure, issues of cue integration directly refer to the structure of
the interface between the sensory system and the perceptual system. In
contrast, issues of conjoint representations refer to structural properties of
the perceptual system itself.
Although we are still far from having a clear theoretical picture of
the kind of primitives that underlie perceptual representations, primitives
33 Conjoint Representations

such as “surfaces,” “objects,”17 or—as temporal analogs to “objects”—


“events” (to be understood as internal, and not as physical concepts) seem
to be among the fundamental pillars of the internal representational
structure of perception. These primitives determine the data format, as
it were, of internal coding. Each primitive has its own proprietary types
of parameters, relations, and transformations that govern its relation to
other primitives. The data structure for the internal representational
primitive “surface,”18 for instance, can be expected to include a set of free
parameters, which refer to attributes such as “color,” “depth,” “texture,”
“orientation,” and so on (again to be understood as internal, and not
as physical attributes) and may also include specific primitive relations
(which may correspond to, e.g., “junctions” and “edges” of various sorts,
“concavities” and “convexities,” “gaps,” or “holes”).
The values of the free parameters, which lie in a specific region of
the corresponding parameter space, have to be determined by the sen-
sory input (and are probably modulated by factors such as “attentional
weight”). The sensory input thus serves a dual function: first, it provides
triggering cues for which primitives are to be activated, and thus selects
among potential data formats in terms of which input properties are to be
exploited. Second, it triggers processes that result in a specification of the
values of the free parameters of the activated representational primitive.
Both aspects have to be dynamically interlocked. On the one hand, values
can only be assigned to free parameters once the data format has been
determined; on the other hand, the activation of a specific data format
requires that the values assigned to the free parameters be in a permissible
range and lie in a specific region of the corresponding parameter space
(if certain types of parameters belong to more than one representational
primitive, their values are likely constrained differently). For example, the
wavelength information in a sensory input appears to be exploited by (at
least) two different types of representational primitives, which we can
tentatively refer to as “surface” and “ambient illumination” (again, to be
understood as internal, not as physical concepts). The two data formats
to which these primitives give rise both include a free parameter for
“color.”19 Accordingly, we have to distinguish different types of “colors,”
depending on the particular primitive to which they belong. Colors that
are attached to the representational primitive “ambient illumination” sub-
serve a different function and exhibit different coding properties than
colors attached to the representational primitive “surface” (cf. Mausfeld
2003). The values of the two different kinds of free parameters, which both
34 Rainer Mausfeld

contribute to the phenomenal attribute of color, are likely to be subject to


different types of constraints.
Although the properties and interdependencies of the free parameters
of representational primitives have to mirror, with respect to the percep-
tual system as an entirety, biologically relevant structural properties of the
external world, empirical evidence strongly suggests that they are codeter-
mined by internal aspects, such as internal functional constraints or inter-
nal architectural constraints, such as legibility requirements at interfaces.
The complex and up-to-now poorly understood interdependencies of free
parameters contribute to the fact that representational primitives defy
definition in terms of a corresponding physical concept (even in the sense
of the latter providing necessary and sufficient conditions for the former);
rather, representational primitives have their own peculiar and yet-to-be
identified relation to the sensory input and may also depend intrinsi-
cally on other representational primitives, in a way that cannot simply be
derived from considerations of external regularities, however appropri-
ately we have chosen our vocabulary for describing the external world.20
In inquiries into the nature of representational primitives, we can and,
taking a specific subsystem of the organism as the unit of analysis, should
actually avoid any notions of the “proper” object of perception and the
“true” antecedents of the sensory input, among the infinite set of potential
causal antecedents (though such notions are, of course, an indispensable
part of both ordinary and metatheoretical discourse). The only physics of
the external world that figures in a formal theory of visual perception is
the physicogeometric properties of the incoming light array. In terms of
these properties, we can completely characterize the relation of represen-
tational primitives to the sensory input, and thus their “proximal seman-
tics,” as it were, which can extensionally be understood as the equivalence
classes of the physical input situations by which they were triggered. The
proximal semantics of the perceptual system is, in other words, defined by
its relation to the sensory system. The proximal semantics (as a purely syn-
tactically defined feature) as well as structural relations among represen-
tational primitives are given by design and are thus essentially impervious
to change by experience.
What is modifiable by experience are the values of certain parameters,
the latitude of which is determined in a highly specific way that is propri-
etary to a structure of perceptual representations. Characteristic exam-
ples are provided by Hans Wallach and Eileen Karsh (1963), who showed
that disparity-related parameters can be recalibrated by the kinetic depth
35 Conjoint Representations

effect, when disparity and motion provide inconsistent shape informa-


tion, and by Joseph Atkins, Jozsef Fiser, and Robert Jacobs (2001), who
showed—with respect to perceiving depth or 3-D shape from 2-D displays
in which disks moved horizontally along the surface of a cylinder and
exhibited corresponding gradients of texture element compression—that
the differential weighting in the integration of visual cues is recalibrated by
the corresponding correlations with haptic cues. In animal ethology, illus-
trative examples are the mechanisms by which birds learn to sing a song
appropriate to their species and region (Marler 1999) or the learning of
the solar ephemeris by bees, as part of a sun compass mechanism (cf.
Gallistel 1998). In these cases, a structure of corresponding representa-
tional primitives has to allow for parameters whose values are based on a
“calibration or checking procedure to insure that the values of those sym-
bols do in fact accurately represent the values of the real world variables
to which they refer” (Gallistel 1998, p. 10). For instance, a sun compass
mechanism “has built into it what is universally true about the sun, no
matter where one is on earth: it is somewhere in the east in the morning
and somewhere in the west in the afternoon. Learning the solar ephemeris
is simply a matter of adjusting the parameters of this universal ephemeris
function so as to make it fit the locally observed motion of the sun”
(Gallistel 2000, p. 1183). In the case of the human perceptual system, the
ontogenetic plasticity, provided by specific structures of representational
primitives, cannot, of course, be understood solely on the basis of physi-
cal considerations of this kind.
With respect to the structural interdependencies of the free parameters
that are potentially involved in a certain input situation, we can, for the
purpose of our discussion, distinguish the case dealing with how different
parameters of a specific type of representational primitive are interlocked
in a certain situation from the more complex case dealing with how
parameters of the same type are interlocked in conjoint representations.
When different aspects of the visual input are exploited by the same
type of representational primitives, for example “surface” representations,
we can encounter situations involving competing interlocked parameters,
say for size and distance,21 orientation and form, or motion direction and
form (which can phenomenally be mirrored in multistable or vague per-
cepts). A change in the value of one type of parameter, say for coding
depth, can, even in cases of otherwise identical stimulus conditions, require
strong changes in other types of parameters, say for coding motion direc-
tion or 3-D form. The demonstration by Erich von Hornbostel (1922)
36 Rainer Mausfeld

using a wireframe cube is a particularly striking classical example showing


that a change in parameters for motion direction—and a concomitant
change in depth parameters—constrains form parameters in a way com-
patible only with nonrigid transformations of form (see also Wallach and
Karsh 1963; Wallach, Weisz, and Adams 1956).
Similar observations have been pervasively made with respect to other
attributes (e.g., Schwartz and Sperling 1983; Dosher, Sperling, and Wurst
1986; Kersten et al. 1992). For instance, motion can codetermine color
in various ways (Hoffman 2003; Nijhawan 1997), and Ken Nakayama,
Shinsuke Shimojo, and Vilayanur Ramachandran observed that “[i]f per-
ceived transparency is triggered, a number of seemingly more elemental
perceptual primitives such as color, contour, and depth can be radically
altered” (1990, p. 497). In many of these cases we do not know yet whether
we are dealing with the problem of how the different free parameters of a
single representational primitive are interlocked, or with the problem of
how representational primitives of the same (or similar) type are inter-
locked. As a rough experimental diagnostic, one might conjecture that
cases in which small changes in a relevant attribute of the input cause rad-
ical changes in other attributes indicate situations in which several repre-
sentational primitives are involved.
Representations that form a conjoint structure are of particular theo-
retical interest in the present context. In the case of conjoint representa-
tions the same aspects of the visual input are simultaneously exploited by
two or more representational primitives of the same type or of different
types, whose parameter spaces overlap.
In sufficiently complex perceptual systems with a high degree of repre-
sentational versatility a given (and sufficiently rich) sensory input is not
likely to elicit only a single representational primitive but rather to trigger
conjoint representations. Conjoint representations require special mecha-
nisms and computational means to handle the interlocked way in which
they exploit the same input.

2.5 CONJOINT REPRESENTATIONS AS A GENERAL STRUCTURAL


PROPERTY OF OUR BASIC COGNITIVE ARCHITECTURE

In picture perception, a physicalistically misconstrued framing of the prob-


lem of the dual nature of pictures highlighted an interesting class of phe-
nomena while it at the same time obstructing the way to theoretically deal
with these phenomena in a fruitful way. Once general properties of such
37 Conjoint Representations

phenomena are explored on a sufficient level of abstraction, it becomes


obvious that cognitive science teems with corresponding phenomena,
which witness, it seems to me, our cognitive capability to simultaneously
handle conjoint representations over the same input. First, I will illustrate
corresponding phenomena with some examples from visual psychophysics,
where one can find a plenitude of corresponding examples indicating that
a given sensory input triggers conjoint representations. Afterward, I will
briefly turn to more complex domains, where I must inevitably resign to
point out, in a more or less allusive manner, some structural similarities
with respect to the issues under scrutiny. For each of these examples, I will
first try to identify potential candidates for representational primitives
that are integrated in an interlocked way with respect to the same input
and then provide, in an unsystematic manner, some observations that
speak in favor of corresponding candidates.

Examples from Visual Psychophysics

Depth We presently do not know much about the representational prim-


itives underlying representations for space and depth. The available evi-
dence suggests that there are, in addition to “surfaces,” probably several
quite different representational primitives whose data format is primarily
determined by some depth-related parameters.22 Representational prim-
itives for local or distant “ambient space,”23 or for dealing with “flat”
spatial situations could be regarded as candidates for these representa-
tional primitives. Three observations may suffice to illustrate correspond-
ing issues.
1. In a picture, we can see an object as partly occluded but still intact.
This can be interpreted as a case of a 2-D surface representation compet-
ing with a shallow 3-D object representation. Even a simple figure-ground
situation, such as Edgar Rubin’s vase, exhibits a very flat 3-D appear-
ance; the “perceived depth separation between figure and ground has not
been well understood,” as Naomi Weisstein and Eva Wong (1986, p. 33)
rightly noted.
2. Even when cues carrying a strong internal weight, such as occlusion,
are violated in a way that is globally incoherent with an interpretation in
terms of a physically distant scene, representational primitives with depth-
related free parameters can still be activated and provide an impression
of space and depth. Reńe Magritte’s 1965 painting Le blanc-seing (de-
picting a lady on a horse within an assemblage of trees) may serve as an
38 Rainer Mausfeld

illustration of this. In this painting, occlusion cues that are globally


incompatible are provided by locally switching foreground and back-
ground, without, however, entirely blocking a depth interpretation
(though the resulting impression is a peculiar one). This applies even more
so for cues that do not seem to carry a strong internal weight, such as
linear perspective or global 3-D incoherence of local depth cues (as in so-
called impossible figures). Also in this case, a shallow 3-D object repre-
sentation may compete with a 2-D surface representation.
3. Often representational primitives of the same type seem to be involved
in conjoint representations within different frames of reference, as it were.
In picture perception, the representational primitive “surface” is involved
both with respect to the perception of the picture surface plane and with
respect to the perception of the scene depicted. In this case, different “sur-
face” representations compete for the depth information that is provided
by the incoming sensory input. In the context of picture perception, the
visual system seems to have a preference for assigning the differential
depth information that is available in the incoming light array to the scene
depicted. If the canvas itself exhibits (suitably chosen) differences in phys-
ical depth, the physical depth signal of the medium, that is, the canvas,
tends to be internally assigned to the scene depicted. Consequently, the
canvas itself looks flat while the picture undergoes corresponding “distor-
tions.” Striking examples are provided by the paintings of Patrick Hughes
(cf. Wade and Hughes 1999).

Color and Brightness As depth-related parameters appear, in a highly


tangled way, in multitudinous internal representations, it is difficult to
identify, in a specific case, the conjoint representations involved. The situ-
ation is less complicated when we are dealing with color- and brightness-
related parameters. Here, phenomenological observations on the interplay
of surfaces and (chromatic) illumination as well as corresponding physical
considerations provide a rich source for theoretical conjectures about
basic conjoint representations. They suggest that the perceptual attributes
“color” and “brightness” are part of the data format of two different but
interlocked representational primitives.24 They can figure as free parame-
ters with respect to the representational primitive “surface” as well as the
representational primitive “ambient illumination.” Both representational
primitives thus form a conjoint representation with respect to the free
parameters “color” and “brightness.” The corresponding regions in the
parameters spaces of these two representational primitives overlap. The
39 Conjoint Representations

visual system then has to provide computational means to deal with sen-
sory inputs that are compatible with different parameter combinations in
this joint region.25
The interplay of the two representational primitives involved is phe-
nomenally mirrored in many peculiarities that are characteristic for color
appearances under (chromatic) illumination. Of particular interest among
these is what Hermann von Helmholtz called seeing two colors “at the
same location of the visual field one behind the other” (1867, p. 407) and
what Karl Bühler referred to as “locating colors in perceptual space one
behind the other” (1922, p. 40; cf. Fuchs 1923a). For instance, in a room
illuminated by a reddish light, we can “see” both the color of the object
(e.g., “white” wall) and the color of the illumination, though there is,
as David Katz observed, a “curious lability of colors under chromatic
illumination” (1911, p. 274). Similar observations hold, with respect to
brightness, for the appearance of surfaces on which a shadow is cast. I will
briefly mention a few other observations that seem to be of relevance for
attempting to understand the internal structure underlying color and
brightness perception.
1. The dual nature of color coding that results from the exploitation of
the input by two different kinds of representational primitives is perceptu-
ally mirrored in what, since Katz’s (1911) groundbreaking work, has been
called “modes of appearance,” in particular a “surface color” mode and
an “aperture or light color” mode. This descriptive taxonomy, which itself
is in need of explanation by some deeper principles, has unfortunately
often been called upon as an explanation itself, thereby confusing the
observation with its explanation.
2. The phenomenal dissociation of brightness and grayness also suggests
different representational primitives in which “brightness” figures as a
parameter. Since Hering, is has been well known that even for achro-
matic colors at least a bidimensional account is necessary, as can be wit-
nessed by appearances such as luminous gray. With respect to painting,
the difference between a “brightish white” and a “whitish bright” is cru-
cial and has been recognized as such since painters became interested in
representing the effects of light (Schöne 1954, p. 203).
3. The Mach card or Ewald Hering’s “stain versus shadow” demonstra-
tion (Fleckschattenversuch) are typical classical phenomena that demon-
strate how certain attributes can modulate the relation between different
representational primitives that exploit a given sensory input. In Hering’s
demonstrations slight changes in figural characteristics of the Alberti
40 Rainer Mausfeld

window, namely, masking of the penumbra of a shadow by a dark line,


are sufficient to induce a switch to a “surface” representation that com-
pletely exhausts the information related to brightness. This is even the case
when the physical construction of the situation—that is, light source, shadow-
casting object, and the process of drawing the boundary—is completely
transparent to the subject. The available perceptual and cognitive “inter-
pretations” are completely overruled by a single geometric characteristic.26
4. The coding properties pertaining to a representational primitive “ambi-
ent illumination” (or transmission medium) resemble, and are probably
related to, coding properties of the “ground” in figure-ground segmenta-
tions (cf. Kaila 1928).

Surfaces and Objects Representational primitives of “surfaces” (cf.


Nakayama, He, and Shimojo 1995) and “objects” seem to be among the
pillars of the internal conceptual structure of the perceptual system.
Hence, it is likely that they themselves are differentiated into families of
corresponding primitives that are intertwined in complex ways. From the
many types of observations that pertain to this issue I will, in an unsys-
tematic way, mention only three examples.
1. There are situations where we can simultaneously see two surfaces at
the same “location” of the visual field. For instance, looking out of a train
window at dusk, we can simultaneously see a red hat on the hat rack and
a green tree at the same location in the window. In psychophysics we can
find, as Franz Faul (1997) did with respect to chromatic transparency,
transitions between transparent and opaque representations of surfaces,
or, as Patrick Cavanagh (1987) found, conditions under which surfaces
are simultaneously opaque and transparent.
2. With respect to conjoint “object” representations striking phenomena
can be encountered in cases where representational primitives that refer to
different levels within some “hierarchy of organization” are involved, as in
the previously mentioned case of so-called impossible figures. An example
of other phenomena that can be interpreted along these line are those,
referred to as object superiority effects, that show that coding properties,
such as threshold, detection, or identification performance, with respect to
elements of the same local stimulus configuration, depend critically upon
whether this element is an essential part of an “object” representation
(Gelb 1921; Lenk 1926; Weisstein and Harris 1974; Gorea and Julesz
1990). On a more abstract level, the process of reading itself seems to be
based on an exploitation of a corresponding capacity.
41 Conjoint Representations

3. The most general class of phenomena that are likely to be caused by


conjoint representations are those traditionally referred to as phenom-
ena of figure-ground ambiguity. The Gestaltists rightly stressed that the
figure-ground organization, which refers to internal, mental aspects, not
to aspects of the sensory input, belongs to the most fundamental aspects
of perception.27 They also observed that different figure-ground organi-
zations for identical sensory inputs would give rise to strong changes in
a variety of other factors such as thresholds or perceptual attributes
(e.g., Fuchs 1923b; Gelb and Granit 1923). Within the framework put
forth here, a change in the figure-ground organization for the same
input is regarded as the phenomenal effect of the working of more basic
principles that refer to the structure and interplay of competing conjoint
representations.

Examples from Other Areas of Cognitive Science


Properties of the architecture of internal representations that have to do
with conjoint representations become even more important when the fac-
ulties in question have to fulfill, on the basis of a given sensory input, more
complex achievements, such as the perception of emotional expressions or
the ability to impute mental states to self and to others. As not much is
presently known about the specific data format of representational prim-
itives underlying visual representations, this situation vastly deteriorates
when we turn to other areas of cognitive science, beyond vision and lan-
guage. I will therefore resign myself to listing a few instances that appear
to me to share, with respect to the issue of conjoint representations, inter-
esting structural similarities. If these similarities are not merely superficial
but rather are grounded in deeper structural properties of our cognitive
architecture—and I think there are good reasons to assume so—their
careful disclosure would facilitate the identification of theoretical issues
that are of great relevance to cognitive science.

Language and Meaning Our ability to internally handle conjoint repre-


sentations appears to be mirrored in various ways in language use. Corre-
sponding conjectures appear to gain some plausibility when we deal with
how we linguistically handle what is provided by our perceptual system.
Many of the examples from visual psychophysics mentioned earlier implic-
itly bear on this issue. A more explicit example is provided by the way in
which we can simultaneously handle deictic-versus-intrinsic, and object-
centered versus viewer-centered frames of reference when we are talking
about spatial relations (cf. Levelt 1984; Jackendoff 1987).
42 Rainer Mausfeld

With respect to language and meaning, observations of “conflicting per-


spectives” abound. Although they are undoubtedly of great theoretical
importance, it is not easy to assess whether and to what extent an analy-
sis in terms of conjoint representations is appropriate in these cases (and,
more generally, how to conceive of the relation between the “internal
semantics” of the perceptual system and the lexical semantics in the Chom-
skyan sense). Because the structural similarities between these two cases
appear striking to me, I will, all the same, briefly mention two of them.
In the semantics of natural languages (in the Chomskyan sense of I-
languages), the internal conditions governing the meaning of words can
encompass simultaneously mutually exclusive aspects of concreteness and
abstractness to which we can refer, in using the word, at the same time.
“The notion book can be used to refer to something that is simultane-
ously abstract and concrete, as in the expression The book that I’m writing
will weigh 5 pounds” (Chomsky 1995, p. 236). It seems to be a pervading
characteristic of natural languages that “a lexical item provides us with a
certain range of perspectives for viewing what we take to be the things in
the world” (Chomsky 2000, p. 36) and that “quite typically, words offer
conflicting perspectives” (ibid., p. 126). For example, “I can paint the door
to the kitchen brown, so it is plainly concrete; but I can walk through the
door to the kitchen, switching figure and ground. The baby can finish the
bottle and break it, switching contents and container with fixed intended
reference” (ibid., p. 128).
These and other observations probably point, or so I believe, to inter-
esting similarities between the structure of the lexicon and the structure of
the representational primitives of the perceptual system. In language, lex-
ical items provide us with a certain perspective for viewing what we take
to be the things in the world; they are “like filters or lenses, providing ways
of looking at things” (ibid., p. 36). The same characterization applies, on
this level of abstraction, to the perceptual system. Its fixed set of repre-
sentational primitives provide, as a perceptual ontology, as it were, a set
of concepts or a perceptual vocabulary by which the signs delivered by
the sensory system are exploited in terms of notions such as “surface,”
“physical object,” “intentional object,” “potential actors,” “self,” “other
person,” or “event” (with respect to a great variety of different cate-
gories and time scales), with their appropriate attributes such as “color,”
“shape,” “depth,” or “emotional state” and their appropriate relations
such as “causation” or “intention.” In this regard, the structure of the
perceptual system seems, in humans, to resemble more the structure of
43 Conjoint Representations

language (more precisely, the structure of the lexicon of I-language) than


the structure of the sensory system.28
The rich conceptual structure of the perceptual system, which extends
far beyond physical aspects of the external world, links the signs provided
by the sensory system to the conceptual structure of language and of other
cognitive systems. The structure of the lexicon, where “notions like actor,
recipient of action, instrument, event, intention, causation and others are
pervasive elements of lexical structure, with their specific properties and
interrelations” (Chomsky 2000, p. 62), seems to partly mirror (and extend)
the conceptual structure of the perceptual system. The representational
primitives of the perceptual system and their structure have, in a cognitive
system as complex as ours, to ensure an appropriate fit of data formats at
the corresponding interface. The property of the lexicon that its items
typically offer conflicting perspectives on what we take to be the things in
the world probably has its counterpart, so I’m inclined to speculate, in the
structural organization of the perceptual system in terms of conjoint rep-
resentations of its representational primitives.
Another, and even more complex, case in point is the use of allegories
in oral or written expositions. Their specific properties, as well as the
ranges within which they can be employed, probably also reflect the capac-
ity to handle conjoint representations, which bear, respectively, on the
relation of medium and message, as it were. Allegories provide a way of
expressing something differently from the literal meanings that are used
(aliud dicitur, aliud demonstratur). For all relevant elements of the expo-
sition they provide two interpretations at the same time, one literal inter-
pretation (sensus litteralis) and one actually intended more abstract
interpretation (sensus allegoricus). One has to understand both at the
same time. The literal meaning serves as a kind of semantic medium to
trigger the intended superordinate meaning. Similar considerations apply
to allegorical meaning in pictorial art, as illustrated in figures 2.3 and 2.4.
Here, the visual input is exploited by representational primitives that
deal with concrete physical objects, as well as by other representational
primitives that deal with the perception of social relations and that read
through this level of representation and exploit the input in more abstract
terms. In the pictures displayed, these abstract terms refer to the “nature”
of social order, as it were.
The allegorical character of figure 2.4 may appear less obvious, because
the objects that figure in the literal interpretation, which refers to people
in the Russian concentration camp Kolyma, belong to the same category
44 Rainer Mausfeld

Figure 2.3
Pieter Brueghel the Elder, The Big Fishes Eat the Small Fishes, 1557, copperplate
engraving, FM 1365, Rijksmuseum, Amsterdam.

Figure 2.4
Gerd Arntz, Kolyma, 1952, linocut, private collection.
45 Conjoint Representations

as the ones that figure in the actually intended interpretation, which refers
to the threatening of the individual by the terror of the state.
The pictures displayed demonstrate again that we cannot understand
the use of allegories solely in terms of representational primitives. Rather,
highly complex interactions of our perceptual faculty and various inter-
pretative faculties are involved, about which we presently know next to
nothing, within the framework of cognitive science.

Pretense Play In pretense play, which is a case of acting as if, in which


the pretender correctly perceives the actual situation, we are also dealing
with a case where the same situation is simultaneously exploited by two
different representational structures. These structures compete, because,
as Alan Leslie put it, “typically the pretense representation contradicts the
primary representation” (1987, p. 415). (This contradiction, however,
always remains, as Johan Huizinga (1938/1986) noted in another context,
constantly fluctuating.) The structural similarity between drama, as a spe-
cial case of pretense play, and the dual nature of painting was emphasized
by Michotte and Michael Polanyi. In the “duplication of space and time
that occurs in theatrical representation,” Michotte noted, “the space of
the scene seems to be the space in which the represented events are actu-
ally taking, or have taken, place and yet it is also continuous with the
space of the theatre itself. Similarly for time also, instants, intervals, and
successions for the spectators belong primarily to the events they are
watching, but they are left nevertheless in their own present. A further
peculiar phenomenon that vividly confirms the unreal character of the
representation concerns the way in which an interval, which really lasts
usually a matter of minutes or seconds, comes by this process of trans-
portation to have the apparent significance of days, months, or even
years” (1960/1991, p. 191 f.). And Polanyi (1970a, p. 231) observed that
“the painting’s self-contradictory flat-depth has its counterpart here in
equally paradoxical stage murders and other such stage scenes. . . . Art
appears to consist, for painting as for drama, in representing a subject
within an artificial framework which contradicts its representational
aspects.”
In infant development spontaneous pretense play emerges at a quite
early stage (at about twelve months), and quickly reaches a state, at about
the age of three, where children are able to engage in complex fantasies
involving imaginary objects, animals, or people. Furthermore, children
are also able to understand the pretense play of other (e.g., Harris and
Kavanaugh 1993).
46 Rainer Mausfeld

An explanatory account of pretense play poses, as Leslie (1987) rightly


observed, deeper puzzles than reality-oriented play, which responds to an
object’s actual properties or expresses knowledge about its conventional
use: “How is it possible for a child to think about a banana as if it were a
telephone, a lump of plastic as if it were alive, etc. If a representational
system is developing, how can its semantic relations tolerate distortions in
these more or less arbitrary ways. Indeed how is it possible that young
children can disregard or distort reality in any way and to any degree at
all? Why does pretending not undermine their representational system
and bring it crashing down?” (Leslie 1987, p. 412). Unlike the cases of the
dual character of pictures or the dual nature of color coding, pretense play
cannot be understood by referring to an “ability to coordinate two pri-
mary representations” (ibid., p. 414) of the same situation. Rather, pre-
tend representations “are in effect not representations of the world but
representations of representations” (ibid., p. 417), which makes pretense
play a case where a primary representational structure (which deals with
how the situation is actually perceived) competes with a superordinate
representation or metarepresentation (which deals with what the pretense
is).

Imputing Mental States to Others and Perspective Taking A perceptual


system by definition serves to couple the organism to biologically relevant
aspects of the external world. For an organism with a mental structure
as rich as ours the relevant aspects of the external world pertain not only
to physical and biological aspects but also to the mental states of others.
The world as we conceive it comprises not only objects and surfaces
with their perceptual attributes but also emotional states and intentions of
others. From an ethological perspective one can reasonably expect that
there is, with respect to the architecture and functioning of the perceptual
system, no fundamental difference between perceiving aspects of the phys-
ical world and aspects of the mental states of others. In either case the sen-
sory input serves as a sign for biologically relevant aspects of the external
world that elicits internal representations on the basis of given represen-
tational primitives.
Evidently and not unexpectedly, the capability to mentally interact with
others is part of the newborn’s biological endowment that quickly
matures to a state where the child can impute mental states to oneself and
to others. The ability to mentally interact with others rests on representa-
tional primitives (whose nature is still at the boundary of scientific eluci-
dation) that have their proprietary ways of exploiting the sensory input. It
47 Conjoint Representations

Figure 2.5
Sergei Eisenstein, Potemkin, 1926, still from the “Odessa steps” sequence, New
York, the Museum of Modern Art, Film Stills Archive.

is an essential characteristic of the way these primitives exploit the sen-


sory input that they go “beyond” those physicogeometrical properties of
the sensory input that are exploited by primitives dealing with the physi-
cal world. They go beyond what may be called physical surface character-
istics of the situation encountered and bear on a more abstract construal
of this situation. We can see the eyes of a person and the direction of that
person’s gaze, and we see a person’s face and simultaneously see that per-
son being angry or fearful. Figures 2.5 and 2.6 may illustrate this, both
with respect to a static 2-D representation of a real face and a drawing
that depicts a culturally shaped abstraction of a face.
From the rich empirical evidence that is available I will only mention
one experiment on complex imitation behavior, by Andrew Meltzoff
(1995). Meltzoff investigated, in a suitably constructed and controlled
experimental setting, whether eighteen-month-old infants interpret “behav-
ior in purely physical terms or whether they too read through the literal
body movements to the underlying goal or intention of the act” (ibid.,
p. 839). In the critical test situation infants were confronted with an adult
48 Rainer Mausfeld

Figure 2.6
Pablo Picasso, Weeping Woman, 1937, etching, aquatint, and drypoint on paper,
Paris, Musée Picasso. © Photo RMN—Gérard Blot.
49 Conjoint Representations

who merely demonstrated the “intention” to act in a certain way, using


entirely unfamiliar objects, but never fulfilled this intention. The adult
tried but failed to perform a specific target act, so the end state was never
reached and thus remained unobserved by the child. For instance, for the
object pair that consisted of a horizontal prong that protruded from a
gray plastic screen and a nylon loop, the experimenter “picked up the
loop, but as he approached the prong, he released it inappropriately so
that it ‘accidentally’ dropped to the table surface each time. First, the loop
was released slightly too far to the left, then too far to the right, and finally
too low, where it fell to the table directly below the prong. The goal state
of draping the nylon loop over the prong was not demonstrated” (ibid.,
p. 841).
The recorded responses of the infants, namely, the number of children
who produced the target act, capitalized on toddler’s natural tendency to
pick up behavior from adults and to reenact and imitate what they see.
Interestingly, “infants were as likely to perform the target act after seeing
the adult ‘trying’ as they were after seeing the real demonstration of the
behavior itself” (ibid., p. 845). They did not reenact what the adult liter-
ally did but rather what he intended to do (they did not produce the tar-
get acts when the physics of the situation, i.e., the movements that are
traced in space, were performed by an inanimate device). The type of
interaction exemplified in this experiment gives rise to a situation that
triggers representational primitives that deal with mental interactions and
with the perception of mental states of others. In such situations a kind of
reading-through with respect to the physical surface characteristics is made
possible, which allows the system to organize these characteristics in terms
of more abstract mental representations.
A similar reading-through with respect to physical surface charac-
teristics underlies mirror self-recognition. This is an achievement that
can only occur when corresponding representational primitives for a self-
representation are available, which are interlocked with those that deal
with the physical surface situation. Whereas most monkey species under
most conditions do not show mirror self-recognition (Tomasello and Call
1997, p. 336), in humans the relevant representational primitives under-
lying mirror self-recognition have matured by the age of about twenty
months.
There are many other cases of highly complex mental achievements,
such as the development of the appearance-reality distinction in young
children (e.g., Flavell, Flavell, and Green 1983), whose structural
50 Rainer Mausfeld

Figure 2.7
Enrico Baj, I funerali dell’anarchico Pinelli, 1972, etching and aquatint, private col-
lection. It depicts the “defenestration” of the Italian anarchist Guiseppe Pinelli
from the Milan police headquarters on December 15, 1969, after the bomb attack
by right-wing extremists at the Piazza Fontana.

properties probably can abstractly be described in terms of (unknown)


conjoint representations, which also may compete for the same input. An
important class of representational primitives that have to be assumed as
an internal representational skeleton for perception are also those that
deal with dynamic situations or temporally organized events with respect
to various types of “objects” (e.g., Zacks and Tversky 2001). With respect
to picture perception this can be illustrated by the etching displayed
in figure 2.7, where the scene depicted is perceived as a single moment
within a sequence of events, and the time slices not depicted are as impor-
tant for what is perceived as the one that is depicted.
Structural similarities between these cases appear to suggest that con-
joint representations and corresponding transformational structures for
properly handling them internally are a fundamental property of our per-
ceptual system. On higher levels of the cognitive system, this property may
have its counterpart in our pervading capacity to simultaneously take
conflicting perspectives in “looking at things and thinking about the
products of our minds” (Chomsky 2000, p. 36). Although an attempt to
fit the aforementioned pieces into the theme of conjoint representations is,
inevitably, already highly speculative, with respect to the perceptual sys-
tem, our general capacity to handle simultaneous, conflicting perspectives
will almost certainly lie beyond the reach of such attempts. As Chomsky
51 Conjoint Representations

put it: “What we take as objects, how we refer to them and describe them,
and the array of properties with which we invest them, depend on their
place on a matrix of human actions, interest, and intent in respects that lie
far outside the potential range of naturalistic inquiry” (2000, p. 21). With
respect to those aspects that we hope are within the reach of naturalistic
inquiry we can, in situations of poor theoretical understanding, only
entrust ourselves to Helmholtz’s guiding principle that “order and coher-
ence, even if they ground on untenable principles, are to be preferred
to the disorder and incoherence of a mere collection of facts” (1867,
p. VI).
Whatever the specific nature of the representational primitives of the
perceptual system turns out to be, their categorical character necessitates,
from a functionalist point of view, additional general mechanisms for
handling continuous transitions in the sensory input, as well as for pro-
viding, whenever appropriate, smooth transitions between internal cate-
gories. Corresponding questions are of particular relevance with respect
to conjoint representations. I will therefore briefly address this issue using
examples from visual perception.

2.6 VAGUENESS, SMOOTH TRANSITIONS BETWEEN


REPRESENTATIONAL PRIMITIVES, AND THE NEED OF A “PROXIMAL
MODE”

The relation of representational primitives to the sensory input has, in an


idealized way, been described in preceding sections in terms of the equiv-
alence classes of physical situations by which they are triggered, or, equiv-
alently, in terms of equivalence classes of output codes of the sensory
system. However, such an idealization is evidently inappropriate, because
it would result in a cognitive architecture with functionally undesirable
properties. Instead, we have to assume, in line with empirical evidence,
that the equivalence classes of physical situations by which representa-
tional primitives are triggered have “fuzzy boundaries,” which, in general,
yield smooth triggering characteristics both with respect to the relation of
a single representational primitive to its triggering class of inputs as well
as with respect to transitions between representational primitives that
exploit the same input. Because triggering a representational primitive is
tantamount to exploiting the sensory input (or the output of the sensory
system) in terms of a specific data format with a specific set of free param-
eters, corresponding “smoothness” requirements apply, as a rule, to the
mappings of physical input features to values of the free parameters.
52 Rainer Mausfeld

Usually, in a given input situation (which can also include dynamic


sequences of inputs) there is a latitude whose extent is determined by the
structure of the joint parameter spaces involved, as to which representa-
tional primitives could be triggered and which values could be assigned to
their free parameters; a latitude that corresponds to an ambiguity about
which of potential external situations could have given rise to the sensory
input. By way of illustration, think, with respect to distance and size, of
the Ames room, or, with respect to surface color and illumination color,
of a white wall under reddish illumination and a reddish wall under white
illumination that both give rise to the same sensory input. In such cases
the visual system often exhibits a preference for some “default interpreta-
tions.” These preferences can be expected to partly mirror different prob-
abilities with which a certain sensory input can be caused, under “normal”
ecological conditions, by different external scenes. However, such ecolog-
ical probabilities do not solely or even predominantly determine “default
interpretations,” as the cases of the Ames room and the Hornbostel dem-
onstration illustrate. Rather, internal constraints that result from various
kinds of stability requirements are, in cases where different combinations
of values can be assigned to the free parameters, likely to play a crucial
role in singling out default interpretations.
Choices between potential values of the free parameters could be made
by a strategy by which global changes, following small variations in the
input (or in the vantage point), in the representational primitives triggered
and in the values of their free parameters are, intuitively speaking, kept
at a minimum, particularly at the interfaces of the perceptual system with
the motorial system and with higher cognitive systems. By such a strat-
egy global stability of superordinate representations could be maintained.
Such a strategy would protect the system from settling, under “impover-
ished” situations, on some definite interpretation that would have to be
changed to an entirely different interpretation following a small variation
in the input. In input situations whose properties are compatible with var-
ious combinations of values of the free parameters (of representational
primitives of the same or of different types), transitions between different
interpretations often appear to be to some extent receptive to modulations
by attentional mechanisms.
Color perception appears to be a particularly conspicuous case of con-
joint representations. Because the same characteristics, with respect to
color or brightness, of a light array reaching the eye can be physically pro-
duced in many different ways (e.g., by either a certain interaction of phys-
ical surfaces and light sources or, using a slide or a CRT [cathode-ray
53 Conjoint Representations

tube] screen, by light sources alone). Representational primitives that sub-


serve different distal interpretations, as it were, compete, on the basis of
relevant cues, for the same input. These different but interlocked repre-
sentational primitives, in which “color” and “brightness” figure as free
parameters, were referred to earlier as “surface” representation and “ambi-
ent illumination” representation. Phenomena related to color and bright-
ness perception provide rich evidence for the way transitions between
representational primitives are handled internally. In the classical lit-
erature, corresponding observations were carefully described and their
importance was properly acknowledged, despite the fact that a suitable
theoretical framework for dealing with them was lacking. Katz, Adhémar
Gelb, Hans Wallach, and many others described a plenitude of situations
in which “very small changes in external stimulus conditions or in inter-
nal modes of perceiving” are accompanied by continuous transitions be-
tween conjoint representations, or—as Gelb put it with respect to color
—between internal states that are “of essentially different nature” (1929,
p. 600). For example, Turhan (1937, p. 46) observed that, under his expe-
rimental conditions, brightness gradients can simultaneously give rise to
two incompatible percepts, one of a curved surface (as would result from
an “interpretation” of the sensory input in terms of a specific nonhomo-
geneous illumination) and another one of a slanted flat surface (as would
result from an “interpretation” of the same sensory input in terms of a
homogeneous illumination). However, the triggering strength of the sen-
sory input does not suffice to tighten an unambiguous “interpretation” in
terms of either of the representational primitives involved. The internal
vagueness with respect to the representational primitives involved is, as
Turhan noted, perceptually mirrored in a peculiar impression of percep-
tual vagueness and indeterminacy.
In color perception, we can deal with the interplay of the conjoint rep-
resentations involved—more specifically, with the relation between the
corresponding free color parameters—in terms of the idealized functional
goals of illumination invariance and scene invariance of the surface color
at a scene location. Because the same sensory input can be compatible
with quite different combinations of values of the free parameters (which
mirrors the different ways in which the input could have been causally
generated) and thus give rise to different functional achievements, the sys-
tem has to guarantee smooth modulations, under small input variations,
of the relations between the representational primitives involved and
thus to provide at least a partial compensation between the relevant
free parameters. A simple observation that witnesses a corresponding
54 Rainer Mausfeld

property is provided by the fact that, for instance, a green light (or a
greenish ambient illumination) and an olive-green surface, whose colors
are yielded by free parameters of different representational primitives,
exhibit some phenomenological similarity, although the classes of appear-
ances these two primitives give rise to could, in principle, have been com-
pletely divorced from each other.
An important consequence of the requirement of ensuring smooth tran-
sitions between conjoint representations is the existence of what is called
a “proximal mode” in perception. The existence of a proximal mode is,
as Rock noted, “not merely of interest as a phenomenological nicety but
rather has important ramifications for a thorough-going theory of per-
ceptual constancy” (1983, p. 254). Evidently, once we have attained the
ability to exercise a suitable “mental attitude,” we can perceptually detach
certain attributes from their “frame of reference” as given by a specific
representational primitive in which these attributes figure. For instance, a
coin lying on the ground at some distance from the observer and being
viewed at a slant is perceived as being circular in shape and of its usual
size. Still, we can also see it, in the proximal mode, as an elliptical shape
of diminutive size. In the same vein, railroad tracks that recede from the
observer toward the horizon are perceived as being parallel. Still, we can
also see them, in the proximal mode, as converging toward the horizon.
Attributes that figure in both types of conjoined representations can,
apparently, be dissociated from aspects that are proprietary to each of the
representations involved. Thus, the existence of a proximal mode helps to
protect the system from adopting a behavior where small continuous
changes in the input result in abrupt changes in internal representations.
It is important to note that only those aspects necessitated by corre-
sponding continuity considerations are accessible to a proximal mode.
There is no proximal mode in the sense of a measurement device miscon-
ception of perception, or in the sense of the (entirely obscure) notion of a
kind of retinal seeing (there is, for instance, no proximal mode for a veridi-
cal seeing of isolated elements of so-called geometrical illusions). Rather,
what can figure in a proximal mode is entirely determined by the structure
of conjoint representations involved.
In color perception, for instance, the proximal mode percept corre-
sponds to that combination of potential values for the free color parame-
ters of both representations involved that is determined by the internal
assumption of a “canonical” or default situation, which, in this case,
would correspond to a spatially homogeneous illumination that does not
55 Conjoint Representations

chromatically deviate from a “normal” one. The small decontextualized


color patches underlying colorimetry are, with respect to the representa-
tional primitives involved, a degenerate situation that is closely related to
the proximal mode. Because such isolated patches proved very useful for
investigations into functions of the sensory system that pertain to color,
they often are misleadingly regarded as the building blocks of color per-
ception. In terms of the representational primitives involved, these iso-
lated patches correspond to in-between stages of internal vagueness—
which is not to be confused with perceptual vagueness (there is no per-
ceptual vagueness in these cases)—where the system has not yet been able
to settle on a data structure in terms of these primitives.
The percept yielded by the proximal mode is sometimes referred to as
the “local color quale.” In many situations, one can focus attention on the
local color quale as such, or on color as a property of surfaces (cf. Arend
and Goldstein 1987); for instance, a spot appearing gray when seen in the
first mode of attention may appear as a shadowed part of a white object
or a illuminated part of a black one in the second mode. Situations like
these, in which it is possible to produce, by slight changes in the mode of
attention, transitions where the “surface gains in whiteness to the same
extent that the illumination looses brightness” are, as Gelb rightly noted,
of “particular theoretical importance” (1929, p. 600). As in picture per-
ception, where we can, with respect to depth, simultaneously have the phe-
nomenal impression of two different types of objects, each of which seems
to thrive in its own autonomous spatial framework, we can also, with
respect to color or brightness, encounter situations where we seem to have
two mutually incompatible representations at the same time, between
which we can, to a certain extent, switch to and fro.
With respect to depth, it is much more difficult than it is with color to
identify representational primitives in which depth figures as a parameter
and that are interlocked to form conjoint representations. For biologi-
cally crucial internal attributes such as “depth,” the corresponding spatial
representations are based on a high redundancy from many subsystems in
order to guarantee a stable representation even in situations where inter-
nal and external conditions deteriorate. In the case of depth, it is particu-
larly difficult to distinguish (1) cases of cue integration with respect to the
same instance of a representational primitive, say, a specific “surface” rep-
resentation, (2) cases in which several representational primitives of the
same type are interlocked or compete (e.g., several “surface” representa-
tions in a transparency situation), or (3) cases in which representational
56 Rainer Mausfeld

primitives of different types are interlocked in conjoint representations.


The requirement of smooth corresponding transitions is, however, of
importance in each of these cases.
Potential candidates for different and probably conjoint representations
in which “depth” figures as a parameter are, on the one hand, those that
deal with the (entirely relative) spatial layout of a scene (e.g., those under-
lying the kinetic depth effect) and, on the other hand, those that deal with
egocentric distance in a full-fledged ambient 3-D space. There also seem
to be specific mechanisms subserving “flat” representations by ignoring
certain aspects of 3-D structure. For many tasks and operations, the avail-
ability of a full-fledged 3-D representation is not necessary or can even be
an impediment. Nevertheless, the visual system cannot simply discard the
corresponding information but has to keep it available internally, because
slight changes in the retinal input might require access to 3-D representa-
tions. As in the case of color, this handling of multiple representations can
phenomenally be either imperceptible or it can be mirrored in multistabil-
ity or in perceptual vagueness. All of these phenomenal accompaniments
can be encountered in picture perception just as much as in other areas of
perception. Picture perception is not special, neither with respect to this
property nor with respect to other perceptual principles. Like all per-
ceptual tasks that involve human artifacts, it rests on and exploits the
complex interactions of given perceptual structures and—most notably in
the case of nonnaturalistic paintings—of various interpretative faculties,
whose properties are at present only poorly understood. Among these
properties is the ability to phenomenally access the different “layers”
of conjoint representations or to exercise attentional control over them,
within the narrow constraints set by the system. Artifacts depend on
human intentions, and their use is therefore subject to interpretation; this
holds for TV screens, microscopes, books, or pictures. One has to under-
stand what they were designed for; nevertheless, they exploit given capac-
ities. From the perspective of the cognitive sciences, picture perception
does not constitute a domain of phenomena bound together by some
domain-specific explanatory principles. Only when we have become aware
that a classification of phenomena in terms of “picture perception” relies
on a pretheoretical commonsense taxonomy can phenomena of picture
perception prove fruitful for directing our theoretical attention to struc-
tural properties of our mental architecture that we otherwise find difficult
to notice because they are an all pervading property of the way we are
designed.
57 Conjoint Representations

ACKNOWLEDGMENT

I should like to thank Franz Faul and Johannes Andres for comments on
an earlier draft of this chapter.

Notes
1. Diels (1922, fragment 12, p. 123).
2. Also, pictorial devices—flattening techniques (Willats, 1997)—have been devel-
oped to control the perceptual balance between the flatness aspect of the picture
and the depth aspect of what is depicted, such as accidental alignments between
two or more parts of a scene and the position of the viewer, the use of mixed and
mutually inconsistent perspectives, or obtrusive surface marks.
3. These phenomena could, from a physicalistic perspective, be described as a
kind of discrepancy between what is physically there, namely, a flat surface,
and the perceptual impression evoked. However, framing the problem this way
amounts to conflating the level of the physical generation process of the sensory
input with the level referring to perceptual mechanisms by which this sensory
input is exploited (cf. Mausfeld 2002).
4. Phenomenological observations that appear particularly salient or enigmatic
do not necessarily have a particular relevance for perception theory. Although
phenomenological observations of various kinds are of prominent heuristic impor-
tance for perception theory, they do not carry a kind of “epistemological superi-
ority.” Phenomenological observations do not provide “direct access” to the
nature of representational primitives; rather, they result from an interplay of vari-
ous faculties, including linguistic and interpretative ones. Thus they are, within a
naturalistic inquiry into the principles of perception, on a par with many other
sources that provide relevant facts and observations.
5. The notion of “representation” is burdened with a high degree of ambiguity
due to its multifarious meanings. Many corresponding locutions in this chapter,
such as pictorial representation, refer to ordinary discourse. With respect to these,
I do not attach much importance, in the present context, to carefully distinguish-
ing different meanings. In the context of explanatory frameworks for certain phe-
nomena, however, I use the notion of “representation,” for example, in the terms
spatial representations or surface representations, to denote elements of postulated
internal structure that are part of an inference to the best explanation. In the con-
text of perception theory, this neither involves particular ontological commitments
nor any reference to the external world. More concretely, a “surface” representa-
tion is not, in any meaningful sense, to be understood as a representation of phys-
ical surfaces. Dispensing with notions of “reference,” “truth,” or “veridicality”
within explanatory frameworks of perception theory is, as Chomsky (1996, 2000)
has argued most convincingly and adamantly with respect to naturalistic inquiries
of the mind, entirely in line with standard methodological principles of the natu-
ral sciences.
58 Rainer Mausfeld

6. Demonstrations for different kinds of situations of conflicting depth cues have


been provided, for example, by Hornbostel (1922), in his striking demonstration
that the image of a rotating wire cube viewed in a mirror undergoes a dynamic
nonrigid transformation (cf. also Adams and Haire 1959); by Ames (Ittelson 1952);
by Epstein (1968), Gillam (1968), Youngs (1976), Rogers and Collett (1989), and
Trueswell and Hayhoe (1993); and by Koenderink, van Doorn, and Kappers
(1994).
7. This is conspicuously illustrated and symbolized by Magritte’s painting La con-
dition humaine, which plays with the many different levels involved.
8. The most extreme versions are trompe-l’oeil paintings that aim to induce
the viewer to perceive the painted object as reality. Such illusionistic effects only
work in cases where the painted 3-D perspective is as shallow as possible, which
explains the highly restricted pictorial themes of trompe-l’oeil paintings (cf.
Mauries 1997).
9. Polanyi (1970a) and Pirenne (1970) introduced the terms “focal” versus “sub-
sidiary awareness” to deal with these observations. Focal awareness refers to the
subject represented, and subsidiary awareness refers to the characteristics of the
surface of a picture.
10. For similar reasons, Michotte (1948/1991, p. 181) emphasized the distinction
between “phenomenal three-dimensionality” and “phenomenal reality.” He
argued that “three-dimensionality and reality are different properties of our per-
ceptions and must be considered as independent dimensions of our visual experi-
ence. . . . By reality we mean an empirical characteristic, the potential for being
manipulated. By three-dimensionality we mean another empirical characteristic,
the capacity for being matched to the volume of a substantial object” (ibid).
11. Even within a single picture, different and globally incoherent local spatial rep-
resentations can be elicited. This was deliberately employed to create certain aes-
thetic effects, as in Piero della Francesca’s painting The Resurrection of Christ (see
Field 1993) or paintings by de Chirico.
12. The action spectra for wavelength-dependent behavior underlying bees’ celes-
tial orientation and navigation depend on more than one pigment, without
exhibiting metameric classes, whereas trichromatic color vision is exclusively
employed in feeding and recognition of the hive.
13. Interestingly enough, Gombrich (1982) also explicitly based his position on
ethological arguments.
14. Michotte was particularly sensitive to the problem of meaning in perceptual
theory, which he regarded as being intrinsic to the structure of primitives that
underlie perceptual organization and that “prefigure” the phenomenal world: “Our
research suggests that this primitive structure occurs in the form of a world of
‘things’ that are separate from one another and are either passive or else ani-
mate bodies ultimately endowed with specific “vital” movements. . . . This possibly
throws some light on the origin of these concepts, but we ought to stress the bio-
logical importance of such spontaneous organization of the phenomenal world
since only such organization could enable the individual (whether human or ani-
59 Conjoint Representations

mal) to adapt its reactions before any individual experience had the opportunity
to provide it with any structure” (Michotte 1954/1991, p. 45).
15. This distinction is different in character from widely made distinctions between
so-called earlier or lower-level systems and higher-level systems. The latter basi-
cally correspond to the sensation-perception distinction as used by Spencer,
James, Wundt or Helmholtz, which refers to an alleged hierarchy of processing
stages by which the sensory input is transformed into “perceptions.” In contrast,
the present distinction refers to two categorically different types of structures and
is more in line with corresponding distinctions by Descartes, Cudworth, or Reid
(cf. Mausfeld 2002, appendix).
16. Computational approaches of the kind pioneered by Marr almost exclusively
deal, with respect to this distinction as I conceive it, with the sensory system; they
have revealed that it has a much richer conceptual structure and greater computa-
tional power than previously assumed.
17. Among representational primitives pertaining to “objects” there are, as corre-
sponding evidence suggests, not only those that pertain to “physical objects” of
various types but also a great variety of specific types that pertain to intentional
physical objects or to biological objects.
18. Again, the internal concept “surface” is assumed to be entirely determined
syntactically, that is, by its data structure and the kind of transformations and
relations that operate on it. It is not, in any meaningful sense, a representation of
physical surfaces. I use the term surface representation only as a convenient abbre-
viation for a postulated representation (whose nature we presently only poorly
understand), whose properties seem to be conveniently describable, at the meta-
theoretical level of the scientist, in terms of perceptual achievements related to
actual surfaces.
19. More precisely, the two different parameters involved can be regarded as
pertaining to the same attribute, if they figure as parameters of the same type in
some superordinate structures and computations. Again, a label such as “color”
serves only as a convenient metatheoretical characterization of a certain type of
parameter.
20. Because of these interdependencies of free parameters, attempts to identify the
representational primitives of the structure of perception and their “data struc-
ture” by investigating attributes like color or depths in isolation are doomed to fail
(apart from lucky coincidences). They are just as futile as it would be to try to
determine a n-dimensional manifold from a random sample of one-dimensional
projections.
21. Even in cases of physically identical input situations, perceptual properties
that have usually been regarded as predominantly mirroring properties of input
channels, such as discrimination, critically depend upon the settings made for con-
joint parameters. A case in point is the observation, known as the Aubert-Förster
phenomenon, that discrimination for objects that subtend the same visual angle is
better when the object is perceived as a small one at near distance than when it is
perceived as a large one at greater distance.
60 Rainer Mausfeld

22. Jan Koenderink argued that “the notion of a depth map as summary repre-
sentation of pictorial relief is hardly tenable.” He concluded that it is likely “that
mental structure contains various (perhaps mutually inconsistent) fragments of
data structures and that only the execution of particular tasks may perhaps draw
on a variety of them and lead to some degree of coordination” (1998, p. 1083).
23. The Ames room demonstration also suggests that an “ambient space” repre-
sentation is triggered according to its own rules and has, in cases in which differ-
ent combinations of values can be assigned to free parameters, its proprietary
“default interpretations,” even if these result in ecologically odd parameter set-
tings for “object” representations pertaining to objects located in this ambient
space. (In the Ames room, a person who walks along the wall opposite the ob-
server, which physically recedes in depth from the observer, appears to shrink in
size.)
24. “Color” presumably also figures as a free parameter in a variety of superordi-
nate primitives that pertain to more complex, biologically relevant aspects of the
external world, such as those pertaining to “edible things” or to “emotional states
of others.”
25. Mausfeld and Andres (2002) found evidence that second-order statistics of
chromatic codes of the incoming light array differentially modulate, by a specific
class of parametrized transformations, the relation of the two kinds of representa-
tional primitives involved.
26. For recent demonstrations that bear on these issues see, for instance, Adelson
(1993), Knill and Kersten (1991), or Buckley, Frisby, and Freeman (1994).
27. Problems of figure-ground segmentations are also a “major obstacle in de-
veloping computational theories,” as Weisstein and Wong (1986, p. 61) noted,
because basic elements that are used in standard computational theories for the
extraction of surface properties are themselves dependent on figure-ground seg-
mentations. Figure-ground segmentation itself is a most fundamental variable that
determines and influences perceptual attributes such as color or depth. Surfaces
that are linked up as “background” can even survive inconsistent disparity infor-
mation, as Belhumeur (1996, p. 342) showed by a stereogram in which we perceive
a continuation of a background object behind foreground strips, even if this is not
consistent with the actual disparity relations for a part of the background section.
28. If, to some interesting extent, this should indeed turn out be the case, com-
paring the functioning of the perceptual system with language, as notably
Descartes and Cudworth did (cf. Mausfeld 2002, appendix), would not merely be
an illustrative or pedagogical metaphor but rather a theory-constitutive metaphor,
which invites exploring “the similarities and analogies between features of the
primary and secondary subject, including features not yet discovered, or not yet
fully understood” (Boyd 1979, p. 363).
Chapter 3
Relating Direct and Indirect Perception of Spatial Layout

H. A. Sedgwick

3.1 DIRECT AND INDIRECT PERCEPTION

Pictures are made and used for many purposes. In my work, and in this
chapter, I focus on just one of these purposes, which is to afford the
accurate perception of the three-dimensional layout of a scene. This pur-
pose may be linked to, and in part derived from, many other purposes—
aesthetic, narrative, historical, architectural, navigational, instructional,
and so on—that give it a social setting and significance, but I shall not
pursue those links here. To say that the purpose is accurate perception
implies that we have knowledge of the true layout of the scene, to which
an observer’s perception of its representation can be compared. The scene
represented may actually have existed at one time, as in a photograph; it
may be pure invention, as in a 3-D computer graphic; or it may be some
combination of invention and reality, as in an architectural drawing. But
it is, in all the cases I am considering, precisely specifiable. I refer to this
represented scene as the virtual scene.
It would go beyond my purpose, and my competence, to pursue in this
chapter philosophical questions concerning what is real and how we know
it to be so. Instead, I shall adopt a naive realism that assumes the exis-
tence of a real world that surrounds us and that we perceive, more or less,
when we look around. Following James J. Gibson, I shall refer to our per-
ception of this real world, or scene, as direct perception. Gibson, and I,
then refer to our perception of a virtual scene that is represented to us in
a picture as indirect perception. These two terms, “direct” and “indirect,”
are to be understood in relation to each other. They refer to two dif-
ferent situations for perception. Direct perception is perception “without
intermediary agents,” to quote a dictionary definition (Webster’s College
Dictionary 1992); it is looking directly at the scene itself rather than look-
ing at a representation of it. Indirect perception refers to looking at a
62 H. A. Sedgwick

virtual scene through the intermediary agency of a representation. Gibson


put it bluntly: “Direct perception is what one gets from seeing Niagara
Falls, say, as distinguished from seeing a picture of it” (Gibson 1979,
p. 147).
Making pictures is, or can be, an art, but making accurate pictures of
three-dimensional scenes is also a technology. This technology has been
developing for a long time and is continuing to develop rapidly, so the
variety of pictorial representations available to us is now immense. Thus
we have not just one kind of indirect perception but rather a multidimen-
sional continuum of kinds of indirect perception.
One source of variation comes from the kind of information that is pro-
vided to specify the represented scene. We may have one or more kinds of
pictorial information, such as perspective, shading, texture gradients, and
so forth, as are found in paintings and photographs; motion information,
as is found in movies and television; stereoscopic information, as in ste-
reograms; or interactive information, as in displays contingent on eye,
head, or body movements or on simulations of any of these.
A pictorial display is itself an object in the real space of the observer. So
a second source of variation is in the kind of information available to
specify the display itself. Such displays are most typically presented on a
flat surface of some sort (although the variations from this are highly
interesting), such as a canvas, a photographic print, a projection screen,
or a computer or television monitor. The information specifying this sur-
face may include the oculomotor adjustments of accommodation and ver-
gence, the stereoscopic disparity of the frame or surface of the display,
motion parallax, and so on. In some situations efforts may be made to
minimize this information, such as fixing the position of the head, limiting
vision to one eye, or requiring the observer to look at the picture through
a pinhole.
A third source of variation is in the interface between the display and
the surrounding environment of the observer. Any representation of a vir-
tual space is always embedded in the real space of the observer. We may
ask, How much of the field of view is taken up by the display—only a few
degrees, as in some computer displays, or a wide angle, as in Cinerama or
IMAX movies? Is the real environment also visible or is it concealed?
Does the display attempt to be immersive (perhaps with haptic feedback
or vestibular stimulation)? To what extent are the exploratory movements
of the observer allowed or inhibited? How casual or controlled is the pres-
entation of the display?
63 Relating Direct and Indirect Perception of Spatial Layout

In one corner of this multidimensional continuum of possibilities, we


have displays that do their technological best to duplicate the perceptual
experience of direct perception. These are the so-called virtual reality dis-
plays. They have many limitations now, but it would be foolhardy to try
to define limits on how close they can eventually come to complete suc-
cess. I think this corner, pushed to whatever limits we imagine may be
possible for it, is a useful reference point in thinking about the relation
between indirect and direct perception.
What would be the opposite corner of this multidimensional contin-
uum? The minimal pictorial display? It is not easy, nor for my purpose is
it important, to pin this corner down, because there are various plausible
candidates. A simple pencil sketch, a patch of texture, the motion of a few
points of light, a few lines seen stereoscopically—each of these is capable
of producing some perception of three-dimensional space. The assertion
that I want to make is that, whichever of these corners we start from, there
is a continuous, seamless path (actually many paths) connecting this cor-
ner of our continuum to the opposite corner.
Thus it seems to me that if we are going to make categorical statements
about pictorial representation and about indirect perception, then these
statements need to be broad enough to include this entire continuum from
the simplest sketch to the most complex virtual reality.
To define direct perception as looking at the real layout of the world
and indirect perception as looking at the virtual space of a representation
leaves open the question of how different these processes are. The most
parsimonious hypothesis is that they are fundamentally the same. Such a
hypothesis is implicit in the research of many vision scientists, who increas-
ingly are using pictorial displays as a convenient way of studying direct
perception, although without always paying sufficient attention to the cur-
rent limitations of such displays, as Jan Koenderink has very eloquently
noted in a recent editorial (1999; also see Hurlbert 1998). Thus we can try
to understand indirect perception and direct perception in the same terms.
Only if this effort fails do we need to postulate perceptual processes or
mechanisms that are unique to indirect perception.

3.2 THE PROBLEM OF DIRECT PERCEPTION

To attempt to understand indirect perception in terms of direct perception


implies that we must first understand direct perception. This is an unfor-
tunate implication in that we are still far from such an understanding. This
64 H. A. Sedgwick

is not just a matter of clearing up the details. There is disagreement, as


well as some lack of clarity, about the most fundamental issues in direct
perception.
For example, we may ask what the most appropriate way is to describe
or characterize what we see when we look at a three-dimensional scene.
There is what I shall call the “standard model,” which goes back at least
as far as Descartes. According to this standard model, what we see in the
first instance is the angular direction and distance from ourselves of every
point in the scene. David Marr referred to such a description as a “viewer-
centered” representation of the scene (Marr 1982). Our more complex per-
ceptions are then built up from this basic information. We determine the
size of a surface, for instance, by geometrically combining its angular
extent (determined by differences in direction) with its distance. The three-
dimensional shape and slant of surfaces are determined in a similar fash-
ion. The perception of angular direction is not regarded by this model as
being much of a problem, because angular direction can be directly re-
lated to the optical projection on the retina, combined with the rotational
posture of the eye in the head. The perception of distance, however, is a
major problem because the projection of a point onto the retina is the
same whatever the point’s distance. Recovering the point’s distance from
its projection thus seems almost an impossibility; it almost seems that any
distance is as good as another, making distance completely ambiguous.
The standard solution is the list of cues to distance—accommodation and
vergence of the eyes, stereopsis, motion parallax, and so forth.
There has been dissatisfaction with the standard model for some time,
but that dissatisfaction has grown more acute and widespread in the past
few decades, and the search for alternatives has been increasingly vigor-
ous and productive (reviewed in Sedgwick 2001). Proposals have been
made, for example, that the geometry of three-dimensional perceptual
space is non-Euclidean, is affine, or is ordinal. Although each of these
descriptive systems may give us some useful angle on the character of
direct perception, it seems unlikely to me that visual space perception by
biological organisms, such as ourselves, adapted to our particular envi-
ronmental niche, is going to be adequately described by any geometry that
is abstract, fixed, and universal.
In what follows, I’ll briefly describe three hypotheses about direct per-
ception, and for each hypothesis I’ll try to describe some of its implica-
tions for our understanding of indirect perception.
65 Relating Direct and Indirect Perception of Spatial Layout

3.3 THE GROUND THEORY OF SPACE PERCEPTION

The first hypothesis, proposed by Gibson in 1950 (Gibson 1950, p. 6), is


that, instead of space per se, what we perceive is the layout of connected
surfaces that compose our visible environment. The foundational surface
of the environment, for terrestrial animals such as ourselves, is the ground
plane. Except for things that float or fly, everything ultimately is sup-
ported by the ground and so is connected through the ground plane to
everything else. Gibson proposed that rather than seeing distance through
empty space we see distance along the ground plane. This conceptual shift
resolves the apparent ambiguity of space perception. Increasing distance
along the ground plane is projected as increasing angular height in the
optical projection, which Gibson called the optic array. This angular
height in the optic array, together with the height of the eye above the
ground plane, thus unambiguously specifies distance along the ground.
Gibson referred to this hypothesis as the “ground theory” of space per-
ception and saw it as a radical departure from the standard model of
space perception.
What Gibson did not know was that the same hypothesis had been pro-
posed, and discussed in some detail, almost 1,000 years earlier by the Arab
philosopher and scientist Alhazen, whose great work on optics only began
to appear in English in 1989 (Alhazen 1039/1989), ten years after Gibson’s
death. Alhazen made the observation that the distances of objects are
uncertain if they are seen protruding above the top of a wall, so that their
bases are not visible, and he argued that distance can only be perceived
between objects that are visibly connected by a surface such as the ground.
I do not yet know the history of this idea during the 600 years between
Alhazen’s death in 1039 and the publication of Descartes Optics in 1637
(Descartes 1637/1965), but I see no trace of it in the standard model,
described earlier, that Descartes advocated.
We must qualify this ground theory hypothesis somewhat because
binocular vision and stereopsis allow us to see the sizes and distances of
objects floating in empty space with some accuracy, especially if they are
fairly close to us. But this is a limited qualification. Research done by
Barbara Gillam and me (Gillam and Sedgwick 1996; Gillam, Sedgwick,
and Cook 1993; Sedgwick and Gillam, submitted), as well as by some
other groups (Glennerster and McKee 1999; Mitchison and Westheimer
1984), shows that if an object floats in front of a background surface,
66 H. A. Sedgwick

rather than in empty space, then stereopsis is used to relate the object to
this background surface rather than to the viewer.
The relation between direct and indirect perception is much more
straightforward from the standpoint of the ground theory than from that
of the standard model. Because the standard model posits distance per-
ception through empty space, from the eye of the observer to the object,
it has traditionally been forced to rely heavily in its theorizing about space
perception on stereopsis and the oculomotor adjustments of the eyes. But
because these are precisely the forms of information absent from paint-
ings, photographs, and other nonstereoscopic pictorial representations,
an explanatory gulf is opened up between the direct and indirect percep-
tion of space. The ground theory, in contrast, is right at home in pictorial
representations of space. As is clear in the first Western work on the sub-
ject, written by Alberti in 1436 (Alberti, 1436/1972), the successful repre-
sentation of spatial relations depends on the accurate construction of the
ground plane. Without a connected layout of three-dimensional surfaces,
pictorial representations tend to collapse onto the picture surface. Some
modern artists, such as Wassily Kandinsky, have done fascinating experi-
ments on the creation of three-dimensional impressions from abstract
objects floating in empty space, but much of the fascination comes from
the lability and ambiguity of these impressions. Recently, Jan Koenderink,
Andrea van Doorn, and their colleagues have performed a series of care-
ful and very interesting experiments on the perception of representations
of sculptural objects floating in empty space. They have found that the
perceptions of such objects are stretched and even sheared in ways that
appear to be essentially arbitrary from one observer to the next (Koen-
derink, van Doorn, and Kappers 1995; Koenderink, van Doorn, Kappers,
and Todd 2000a).
Our predisposition to perceive the environment as a layout of con-
nected surfaces underlies the effectiveness of simple presentations such as
line drawings. A horizontal line across the middle of the page is sometimes
enough to invoke the perception of the ground plane. A complex archi-
tectural space of connected surfaces can be evoked by a few converging
and parallel lines.
In complex scenes, not everything rests directly on the ground or floor.
Instead, a lamp may rest on the table, or a book may rest on a shelf
attached to the wall. But the wall extends down to the floor, as does the
table, so that ultimately everything that does not float or fly is supported
by the ground. To apply the ground theory of space perception to com-
67 Relating Direct and Indirect Perception of Spatial Layout

plex environments, it must be generalized to include such scenes. For


some time I have been interested in what I refer to as the “nested hierar-
chy of contact relations,” which establishes the spatial layout of scenes
through complex sets of connected surfaces (Sedgwick 1987a; Sedgwick
1987b; Sedgwick 1989).
Jeanette Meng and I have investigated observers’ ability to perceive
spatial relations in such situations (Meng and Sedgwick 2001). Using
computer graphics, we asked observers to move a marker along a track on
the ground until its distance to the observer matched that of a cube rest-
ing on a block that in turn rested on the ground (figure 3.1). Holding
the cube and the top of the block constant, we varied the block’s point of
optical contact with the ground. In this way we progressively increased the
perceived distance of the block. The perceived distance of the cube also
increased appropriately, showing that its perceived location relative to the
ground was effectively mediated by the block on which it rested.
In another experiment we varied the height of the block and with it the
height of the cube in the image plane. If observers were influenced simply
by the “cue” of height in the image plane, this should have increased the
perceived distance of the cube, but it did not. Because the base of the
block, where it made contact with the ground, did not change, the per-
ceived distance of the cube did not change either. The observers’ distance
settings did become more variable, however, reflecting the increased sepa-
ration between the ground and the cube’s surface of support.
Recently (Meng and Sedgwick 2002) we have looked at a somewhat
more complex situation, in which the cube rests on one block and the
marker and its track rest on another. When the two blocks are the same
height, the variability of distance settings is reduced, suggesting that
observers establish a shortcut directly between the tops of the two blocks,
based on the coplanarity of the two surfaces, rather than relating the cube
and marker through the ground. These exploratory experiments are just a
beginning. If we wish to apply Alhazen’s and Gibson’s ground theory to
the perception of complex spatial layouts, either real or virtual, much
more work remains to be done.

3.4 ENVIRONMENT-CENTERED PERCEPTION OF SPATIAL LAYOUT

The second hypothesis I wish to discuss holds that direct perception is pri-
marily environment centered (Sedgwick 1983). By “environment centered”
I mean that what is most salient in our perception of spatial layout is how
68 H. A. Sedgwick

Figures 3.1a and b


In this experiment using computer graphics, observers moved a marker along the
track on the ground until it appeared to be at the same distance as the small cube
resting on top of the block. In the two conditions shown here, the positions of the
cube and of the top of the block were held constant in the image while the block’s
pictured contact with the ground was varied by decreasing (from a to b) the ver-
tical dimension of the block. The perceived distance of the cube was effectively
mediated by the changing distance of the block, as specified by the location of the
bottom of the block relative to the ground. Reproduced and modified with per-
mission from Meng and Sedgwick 2001; original displays were in color.
69 Relating Direct and Indirect Perception of Spatial Layout

objects’ surfaces are related to the principal surfaces of the environment.


In contrast, the standard model is viewer centered, meaning that spatial
layout is perceived primarily in relation to our own viewpoint.
Let us first consider how this distinction applies to the problem of
size and distance. In the standard, viewer-centered model, the perceived
size of an object is based on its perceived distance from the observer. This
implies that perceived size and perceived distance should vary together;
for example, errors in perceived distance should be reflected in errors in
perceived size. This idea, dubbed the “size-distance invariance hypothesis”
(Kilpatrick and Ittleson 1953), has been examined empirically many times
and is not well supported by the evidence (for reviews, see Gillam 1995;
Sedgwick 1986).
In the environment-centered model the perception of size is based on
the object’s relation to the context provided by the environment. Gibson
(1950) provided an example of such contextual information for size,
pointing out that surfaces such as the ground are usually textured, with
the elements of this texture being uniform, at least statistically, across the
extent of the surface. The relative sizes of objects resting at various loca-
tions on such a textured surface are specified by the number of texture
elements that each object covers. The objects’ distances from the viewer
are irrelevant to this relationship. In this view, Gibson pointed out, the
perceived size of an object is a by-product of the perceived scale of the
environment. As Gibson says, “the size of any particular object is given by
the scale of the background at the point to which it is attached” (Gibson
1950, p. 181; emphasis in original). In addition to texture scale, the scale
of the environment is also specified by perspective. For example, on an
extended ground plane, the line of sight to the horizon is parallel to the
ground and so intersects objects resting on the ground at a constant height
(equal to the height of the observer’s eye). Thus, the horizon sets a con-
stant scale across the entire ground plane. The relative height of each
object is specified by what I refer to as the “horizon ratio relation”: the
ratio of the object’s total optic array projection divided by the projected
portion of the object below the horizon (Sedgwick 1973). As with texture
scale, this ratio specifies the relative sizes of objects independently of their
distances from the viewer.
In the environment-centered model, distance from the viewer’s eye has
no special significance. Distance is perceived between objects and, as dis-
cussed earlier, is perceived along surfaces rather than through empty space.
More fundamentally, though, in the environment-centered model distance
70 H. A. Sedgwick

loses its primacy altogether. What is primary instead is location, the point
at which an object contacts a supporting surface. Given the set of contact
relations between the objects and surfaces of the environment, distances
between locations can be derived as needed (Sedgwick 1983, 1987b).
Finally, let me illustrate this distinction between environment-centered
and viewer-centered models with the example of surface orientation. As
Gibson and Cornsweet (1952) pointed out, we can define the slant of a
surface either relative to the observer’s line of sight or relative to an envi-
ronmental surface such as the ground plane. They referred to these as
“optical slant” and “geographical slant,” respectively. The properties of
these two types of slant differ in important ways. Consider a flat surface
slanted relative to the ground, say at an angle of 45°. The geographical
(or environment-centered) slant of this surface is a constant 45° along its
entire length. It also remains a constant 45° as the observer moves around
in the environment. In contrast, the optical (or viewer-centered) slant,
where the line of sight meets the surface, changes as the line of sight
sweeps along the surface or as the observer moves around.
Surfaces that have the same environment-centered slant are parallel
to each other and thus share the same vanishing line in perspective. Each
distinct vanishing line in the optic array thus corresponds to a family
of parallel surfaces and specifies their environment-centered orientation
(Sedgwick 1983).
Marr suggested that the visual system first finds the viewer-centered
structure of the environment and then can derive other information, such
as parallelism, from it (Marr 1982). Some years ago Steve Levy and I
tested this idea, using computer graphics images, by asking observers to
adjust the orientation of one surface until it matched either the viewer-
centered or the environment-centered slant of another surface (Sedgwick
and Levy 1985). We found that observers were more precise in making
environment-centered matches. This makes it seem more plausible that
environment-centered slant is primary than that it is derived from viewer-
centered slant.
What are the implications of the environment-centered model for the
relation between direct and indirect perception? It seems to me that it
offers us some understanding of the ease with which a representation can
embed a virtual space in the real space of the observer. An environment-
centered virtual space can be coherent within itself without necessarily
being clearly related to, or even commensurable with, the real space of the
observer. Virtual objects’ sizes relative to the scale of their virtual envi-
71 Relating Direct and Indirect Perception of Spatial Layout

ronment, their locations relative to the layout of virtual surfaces, and the
slants of their surfaces relative to the surfaces of this virtual space may all
be more salient to indirect perception than these virtual objects’ viewer-
centered distances or the sizes or slants that could be derived from such
distances.1

3.5 CROSS-TALK

My third hypothesis concerning direct perception is that the perceived


three-dimensional spatial layout of a scene is influenced by cross-talk from
the scene’s projection in the optic array (Sedgwick 1991).2
The recent literature on the perception of distance over the ground
offers a good example of what I am calling cross-talk. Suppose that we
scatter a number of objects over an open field and then we ask an ob-
server to estimate distances between pairs of these objects. From the view-
point of the observer, the projection of the ground plane is compressed, so
the angular extent of a distance radial to the observer is considerably
smaller than the angular extent of an equal distance in the frontal plane of
the observer. Charles Levin and Ralph Haber (1993) have shown that this
difference in angular extent influences observers’ perception of distance,
resulting in an underestimation of radial distances relative to frontal dis-
tances. The perceived relative distances are a compromise between the
true relative distances and their optic array projections. What is critical
here is not distance per se but the relation between distance and its pro-
jection. Jack Loomis and John Philbeck (1999) have shown that with
monocular viewing the effect of compression is invariant when all the
distances are scaled up while the same compressive relations are main-
tained by increasing the eye height proportionally. In all these situations
there is good information available for the true distances, but the per-
ceived distances are nevertheless affected by cross-talk from their optical
projections.
In direct perception, the effects of cross-talk from the optical projection
appear for the most part to be relatively small. When we turn to indirect
perception, however, we are introducing a projection surface, such as the
surface of a photograph or painting, and the perception of this projection
surface, I believe, gives greater salience to the projective relationships,
thus increasing the strength of the cross-talk.
Let us consider one example from a study that I did with Andrea
Nicholls (Sedgwick and Nicholls 1993a). We used monocularly viewed
72 H. A. Sedgwick

computer-generated scenes of a rectangle on the wall of a room. The wall


was slanted so that the projection of the rectangle was compressed on
the surface of the picture. We asked observers for perceptual estimates of
the width-to-height proportion of the rectangle in the scene. We held that
proportion constant on average over a series of trials but varied the pic-
tured slant of the rectangle, so that the width-to-height proportion of its
projection on the picture surface became increasingly compressed. As we
increased the compression on the picture surface, observers’ estimates of
the rectangle in the scene became more compressed. This is a clear exam-
ple of cross-talk from the picture surface to the perception of the virtual
scene.
In another experiment (Sedgwick et al. 1991), we similarly found cross-
talk in the indirect perception of virtual size, comparing the relative size of
two rectangles. We also found that when we increased the salience of the
information for the surface of the picture, by adding a frame and surface
texture, we increased the amount of cross-talk from the picture surface
to the virtual scene. A related result has been found by David Eby and
Myron Braunstein (1995), who showed that adding a frame to a picture
diminished the perceived depth within the pictured scene. Koenderink and
his colleagues have shown that viewing a picture with both eyes, which
provides additional information for the picture surface, rather than with
just one eye, also has a flattening effect on the virtual object (Koenderink,
van Doorn, and Kappers 1995; Koenderink et al. 2000a).
The hypothesis of cross-talk from the picture surface to the pictured
scene helps to resolve an important issue in our understanding of indi-
rect perception. This is the problem of what happens to perceived virtual
space when we view a picture from the wrong viewpoint. An essential
part of creating an accurate representation of a three-dimensional scene
is to choose the viewpoint, the place from which the eye or the cam-
era views the scene (Sedgwick 1980). When the representation, a photo-
graphic print, for example, has been created, there is now a point in the
real space in front of the photographic print where the observer’s eye must
be placed to obtain the same view that the camera had. If the eye views
that representation from some other viewpoint, the perspective of the
photograph now specifies a different virtual space, one that is a distortion
of the original virtual space. For example, if the viewpoint is displaced lat-
erally, then the virtual scene specified by perspective is sheared. If the
viewpoint is too near or too far, then the virtual space is compressed or
expanded, respectively (Sedgwick 1991).
73 Relating Direct and Indirect Perception of Spatial Layout

The perceptual problem is that observers tend not to notice these dis-
tortions. Someone walking through a photography exhibit tends not to
notice distortions of the virtual space of the pictures, even though the
point of observation is continually changing. This has led to the sugges-
tion that there is a perceptual compensation mechanism (Pirenne 1970,
p. 162). Such a mechanism is hypothesized to recognize that the observer
is looking at the representation from the wrong viewpoint and to inter-
nally correct for the distortion that this produces. The existence of such a
mechanism in indirect perception would, it seems, open up a considerable
gulf between indirect and direct perception. There can be no need for such
a compensation mechanism in direct perception, because the viewpoint of
the observer is always, by definition, the correct viewpoint. Thus this
hypothesized compensation mechanism, which would need to be a mech-
anism of formidable complexity to correctly solve the problem of com-
pensation, would exist for indirect perception alone.
The cross-talk hypothesis offers an alternative to the compensation
hypothesis. As the observer moves around, the virtual scene specified by
perspective distorts, but the surface of the picture does not change. Thus
cross-talk from the picture surface to the perceived scene will tend to have
a conservative effect, that is, to reduce the perceived distortions of vir-
tual space. Unlike the compensation hypothesis, however, the cross-talk
hypothesis does not predict that there will no perceived distortion. It only
predicts that the amount of perceived distortion will be less than what is
specified by perspective.
I shall briefly describe one study, done with Nicholls (Sedgwick and
Nicholls 1994), that addresses this issue. Using computer-generated scenes
with slanted rectangles similar to those I described earlier, we looked at
the effect of reducing the size of the picture, which is optically equivalent
to viewing the picture from a distance farther than the correct viewing dis-
tance. The depth dimension of the virtual space specified by perspective is
thus stretched, increasing the optically specified slant and width-to-height
proportions of the rectangle on the wall. We found that the perceived pro-
portions of the rectangle did indeed increase, but not by nearly as much as
perspective predicted. We were able to account for most of this shortfall,
however, with a control condition that measured the cross-talk from
the proportions, unchanged by minification, of the projected rectangle on
the surface of the picture. When we also included controls for the effects
of the smaller angular size of the minified picture and the smaller angular
size of the minified rectangle within it, both of which tended to reduce
74 H. A. Sedgwick

perceived slant, we were able to account for almost all of the discrepancy
between the perceived virtual space and the optically specified virtual
space. We were able to do this without invoking any special perceptual
mechanism to compensate for being at the wrong viewpoint. This, along
with a number of other results (Rogers 1995; Sedgwick and Nicholls 1996;
Sedgwick, Nicholls, and Brehaut 1995; Yang and Kubovy 1999), suggests
to me that the hypothesis of a special compensation mechanism in indirect
perception is unnecessary and that we can account for the indirect per-
ception of virtual space, even when seen from the wrong viewpoint, in
terms of the same mechanisms, such as cross-talk, that we find operating
in direct perception.
Let me briefly sum up. The technology of accurate spatial displays
produces an ever-changing, multidimensional continuum of pictorial rep-
resentations, ranging from the most compelling virtual realities to the sim-
plest pencil sketches.
We may imagine indirect and direct perception as a lock and key. When
we have the right theory of direct perception, it may provide the key that
opens up our understanding of the whole continuum of indirect percep-
tion. I have described three hypotheses about direct perception that may
provide a good fit with the problems of indirect perception.
First, direct perception is attuned to the complex spatial layouts of con-
nected surfaces that compose much of our environment. Display tech-
nologies have for a long time been well suited to the representation of such
layouts.
Second, direct perception is environment centered—attuned to geo-
graphical orientations of surfaces, locations of objects on those surfaces,
and sizes relative to the scale of the environment. This facilitates the indi-
rect perception of self-contained virtual scenes embedded within the real
space of the observer.
Third and finally, there is a dual awareness in direct perception—of the
three-dimensional scene and of its optical projection—and the perception
of each is influenced by cross-talk from the other. In indirect perception
there is a corresponding duality between the perception of the virtual
scene and the perception of the picture surface. Cross-talk from the stable
picture surface tends to reduce the perception of distortions in the virtual
scene when the picture is viewed from the wrong viewpoint.

Notes
1. This is not to deny the possibility of a viewer-centered mode of perception,
especially in near space, and in some visual-motor activities, although such situa-
75 Relating Direct and Indirect Perception of Spatial Layout

tions can easily be mistaken as viewer centered when they are not. Instrumental
vision is rarely exactly viewer centered, where the center is taken to be the nodal
point of the eye, or the cyclopean eye of binocular vision. For manual tasks the
important relations are between the hands and the object; for locomotor tasks
they are between the feet and the goal; for tool-using or vehicular tasks they are
between the tool and its object or the vehicle and its goal; and so forth. It may be
that such tasks are better understood in environment-centered terms.
2. There is also cross-talk in the other direction: if we try to attend to the scene’s
projection, our perception is strongly influenced by the three-dimensional scene.
In indirect perception, cross-talk from the pictured scene to the picture surface can
account for some geometrical illusions, such as the Ponzo illusion (Sedgwick and
Nicholls 1993b).
This page intentionally left blank
Chapter 4
The Dual Nature of Picture Perception: A Challenge to Current
General Accounts of Visual Perception

Reinhard Niederée and Dieter Heyer

4.1 INTRODUCTION

Pictures form a significant part not only of our visually oriented culture
but also of the stimulus material employed by perceptual psychologists.
Nevertheless, with few exceptions (e.g., Rock 1984), standard textbooks
on the psychology of perception tend to ignore the challenging pecu-
liarities of picture perception, and the same goes for mainstream vision
science.
One major reason for this is that, first of all, attention is restricted to
realistic pictures in the strict sense, based on central perspective, such as
Renaissance paintings or photographs. In a second step, then, the percep-
tion of these pictures is conceived as being on a par with the perception of
a scene as seen through a window. Hence, on this account, the perception
of a picture, that is, the perceptual emergence of pictorial space in the
observer’s consciousness, need not be distinguished in principle from the
perception of an everyday scene. Such a conception of picture perception
implies that the spatial percept evoked by a flat realistic picture is simply
a perceptual illusion, which is not different, in principle, from other well-
known perceptual illusions such as the Ames room.
Except for limiting cases such as perfect trompe l’oeil, this is not the
case, however. That is, as a rule, nobody misperceives a painted scene as
a real one. Rather we speak, without hesitation, of seeing a painting of
a scene. As Edmund Husserl (1980), Michael Polanyi (1970b), Maurice
Pirenne (1970), James J. Gibson (e.g., 1979), and others have forcefully
pointed out, this is due to what is often called the dual nature of picture
perception: the evoked percept involves both the perceived surface of the
picture as part of perceived real space and the experience of pictorial
space, both aspects being subtly interwoven. If our visual system were able
78 Reinhard Niederée and Dieter Heyer

to capture merely one of these aspects, then picture perception as we know


it, and along with it the cultural achievements based on it, could not exist.
For, if the pictorial surface were invisible, pictures would cause illusions
in the strict sense and thus would be difficult to handle as objects in every-
day life. Conversely, if just the pictorial surface were visible, then pictures
would reduce either to pieces of decoration or to mere carriers of conven-
tional signs.
Contrary to the window analogy mentioned earlier, some authors have
in fact argued that it is only the flat material picture object that is per-
ceived. The emergence of pictorial space in the observer’s consciousness is
then counted as a phenomenon fundamentally different from (“direct”)
perception. For instance, in his early phenomenological writings, Jean-
Paul Sartre (e.g., 1940) maintains that the experience of pictorial space
is a matter of imagination based on the perception of the material picto-
rial surface (cf. Wiesing 2000, p. 43 ff.). Another widespread move of this
kind is to assume a two-stage process, the first stage of which corresponds
to the perception of the pictorial surface. This percept is then conceived
as being the basis for an intrinsically different additional interpretation
process in a semiotic sense, comparable to reading (“reading a picture”).
In this chapter, in contrast, a unifying psychological perspective will be
taken that understands both aspects of the dual nature of picture percep-
tion as genuine and interrelated features of a complex perceptual process,
which is to say that both aspects are assumed to rely on essentially the
same set of basic perceptual mechanisms. So far as this general assump-
tion is concerned, we thus follow the path already taken by other psy-
chologists of perception, such as Gibson (1979), to mention only one.
To preclude misunderstanding, we would like to add that despite the
emphasis placed here on genuinely perceptual aspects, we do not, of course,
mean to deny the potential relevance of imagination proper or interpre-
tation (in a semiotic sense) in the context of picture perception. These
concepts should not simply be lumped together, however, with the notion
of perception, even though it must be conceded that the boundary between
these concepts seems inherently fuzzy. Furthermore, perceived objects
other than pictures may equally well trigger processes of imagination or
serve as (more or less conventional) signs in a semiotic sense. These issues
will therefore not be discussed here.
From the perspective of visual science, the dual nature of picture per-
ception might at first sight appear to be a marginal, if not an anomalous
phenomenon, dependent on utterly artificial stimuli. Even then, vision sci-
79 The Dual Nature of Picture Perception

ence would have to meet the challenge posed by this phenomenon. For
reasons to be pointed out, we are convinced, however, that phenomena
of that kind should not be considered marginal. On the contrary, we
believe that general theories of perception will profit substantially from
taking such complexities into account from the start. This would require,
however, a corresponding modification of standard scene-based overall
accounts of perception.
From this very perspective, pictures might indeed turn out to be natu-
ral objects of study not only for students of picture perception but for
visual scientists in general. To simplify matters, we will restrict attention
mainly to realist pictures (central perspective). Mutatis mutandis, our
considerations are meant to apply to other cases as well.

4.2 WHERE TO “LOCATE” THE DUAL NATURE OF PICTURE


PERCEPTION: A ROUGH SKETCH

As the reader will have noticed, we have been talking about the dual
nature of picture perception and not about a “dual nature” of pictures
themselves. That is, duality is considered here as a feature of an observer’s
percept caused by a pictorial stimulus—the perceptual experience “of a
picture.” This duality has its origins in the visual processes underlying the
percept. Trivially enough, we will fail to find any such duality in the
“world out there,” that is, in the physical stimulus itself. Nor would it do
justice to the phenomenon at issue just to point at the “discrepancy”
between the real physical stimulus, a flat material object, on the one hand,
and the “illusory,” purely mental, pictorial space evoked by it, on the
other.

The Current Standard Framework of Perception


To spell out more clearly what is meant by this distinction, let us first give
a rough and simplified sketch of the current standard framework of per-
ception. Its basic ingredients are the following. An observer is facing a
physical scene, that is, a spatial arrangement of material objects, illumi-
nated by some light source(s), the so-called distal stimulus. The light is
partially reflected (or transmitted) by the objects (surfaces) in the scene.
The light pattern reaching the eye acts as a medium between the scene and
the observer’s eye, where, via the eye’s lens, it projects onto the retina cre-
ating the “retinal image” (proximal stimulus), or, in the standard case of
binocular vision, a pair of retinal images. The two-dimensional retinal
80 Reinhard Niederée and Dieter Heyer

image itself must not, of course, be confused with “what is seen” by the
observer. Rather, by means of a stimulation of a field of retinal recep-
tors (rods, cones) it serves only as an “input” (“raw data”) to a complex
visual process. In the course of this process the visual system extracts
and integrates cues (or, in Gibsonian parlance, picks up invariants) that
eventually yield a richly structured conscious “output,” the percept. At
this level, meaningful “wholes” in the sense of Gestalt psychology (e.g.,
perceived objects), endowed with (perceived) shape, color, size, spatial
position, orientation relative to the observer, motion, and so on, are orga-
nized into a dynamic, more or less coherent, perceived scene. Note that
although the percept is part of our conscious experience, the underlying
visual processes (including the relevant cues) for the most part are not
accessible to consciousness. Needless to say, our present understanding of
these processes still is rather fragmentary.
In sum, this scheme assumes that a physical input evokes a conscious
experience, the percept. From a functional viewpoint, the latter in turn is
usually conceived as a partial 3-D scene-representation, whose compo-
nents may or may not be “veridical” with respect to the underlying phys-
ical scene. As usual in vision science, the term “perception” is employed
here, regardless of whether or not veridicality is met; that is, the evoca-
tion of a visual illusion by an external stimulus is equally counted as an
instance of perception. From a Helmholtzian viewpoint, this process
would be described as an unconscious inference yielding the most likely
hypothesis about a state of affairs “out there” that could have caused the
retinal input.
From the perspective of the scheme just outlined, it is indeed fairly
obvious that the notion of duality of picture perception pertains to the
visual processes and their “output,” that is, to the percept. However, in
being centered around the concept of a scene-representation, this scheme,
strictly speaking, does not allow for the possibility of this kind of duality.

The Meaning of “Duality”


Unlike the simultaneous awareness of the perceived size and perceived dis-
tance of an object, of perceived incident illumination and perceived sur-
face color (e.g., Katz 1935), or of a perceived transparent layer and the
perceived object seen “through” it, which in principle fit into the frame of
a single consistent scene description, a duality of the kind under consider-
ation does not. For, how could a flat opaque surface on a wall and a trans-
parent opening in that wall with a scene being visible through it coexist at
81 The Dual Nature of Picture Perception

the same location? On the preceding account, however, the system should
always come up with a single (most “simple” or “likely”) scene descrip-
tion. That is, we should either have (a more or less “veridical”) percep-
tual impression of a surface as part of perceived real space or the illusory
impression of pictorial space seen through a windowlike opening. (Here
and in what follows the term “impression” refers to perceptual experiences
and does not necessarily imply vagueness or the like.) At best, bistability—
perceptual switches between the two percepts—might occur, as known
from the well-known Necker cube and many other examples. Regarding
picture perception, Ernst Gombrich (1960) in fact advocated such a view,
insofar as he postulated that we could only attend to one of the two just-
mentioned aspects at a time, ruling out a simultaneous coexistence of both
perceptual impressions.
Against this position, a number of authors—rightly, we believe—have
argued that in many situations we are in fact aware of both aspects simul-
taneously, this double impression making up what we call a prototypical
perceptual experience of a picture (e.g., Polanyi 1970b; Pirenne 1970).
What may be shifted is attention. Indeed, we are hardly ever simultane-
ously aware of both aspects in a full-fledged manner, indeed. In many
situations, pictorial space is the more salient aspect, but our actions con-
cerning a picture will require an awareness of the picture as a flat object in
real space. In fact, even if attention is largely concentrated on one aspect,
there often is at least a residual awareness of the other aspect, too. The
dynamics of attention certainly needs to be taken into account in this con-
nection, but this is true of perception in general and does not imply bista-
bility as it is experienced, for instance, in the case of the Necker cube.
For brevity, the dual aspects of picture perception at issue will hence-
forth be called the planar and the spatial aspects of a picture percept,
respectively, even though the planar aspect itself does of course also
include a spatial component with respect to perceived real space. It is
worth noting that the distinction between planar and spatial aspects of a
picture percept involves not only genuinely spatial features but at the
same time a duality of aspects pertaining to perceived color. For instance,
what looks like a pattern of black, gray, and white patches on the picture
surface (planar aspect) may go hand in hand with the perceptual impres-
sion of a scene composed of objects of different achromatic colors under
certain perceived lighting conditions. As the case may be, this may include
perceived shadows, perceived light sources, perceived translucency, or
perceived gloss. Of course, the same goes for chromatic aspects of color
82 Reinhard Niederée and Dieter Heyer

vision. To our knowledge, this aspect of the duality of picture perception


has been largely neglected in the relevant literature (but see, for instance,
Schöne 1954, appendix B).
Most important, the spatial and the planar aspects do not simply coex-
ist in parallel but are related to each other. They are experienced as a per-
ceptual unit. Whenever we speak here of one’s being aware of both aspects
“simultaneously” or “at the same time,” this kind of unity is tacitly implied.
For this kind of phenomenal binding, too, suitable concepts are missing in
the orthodox scheme of perception sketched earlier. Moreover, in most
situations there is a notable perceptual asymmetry between the planar and
the spatial aspect insofar as only the former is experienced as belonging to
perceived “real space.” To spell this out, perceptual degrees of “looking
real” (cf. Michotte 1948, 1960) and corresponding perceptual hierarchies
need to be taken into account, too.
Not all objects called pictures in our culture evoke dual percepts of
the kind just discussed. For instance, in Andrea Pozzos’s famous ceiling
The Transmission of the Divine Spirit (1688–1694, Chiesa di Sant’Ignazio,
Rome; cf., e.g., Pirenne 1970), the painted pillars appear spatially real
from a specific viewing position. That is, only the spatial aspect is present
in the percept. This is an instance of a trompe l’oeil in the strict sense,
where a picture is not seen as a picture anymore in that there is no simul-
taneous awareness of planar and spatial aspects. Note that in our context
one had better not speak of a painting itself as a trompe l’oeil but, rather,
of a trompe l’oeil effect. For we are speaking of a perceptual phenomenon
that is also dependent to some extent on viewing conditions, such as the
position of the observer. Similarly, for instance, a trompe l’oeil effect
can sometimes occur if parts of an ordinary realistic painting are viewed
monocularly through a reduction tube, or in stereoscopic vision, where
two slightly different pictures are presented to the right and the left eye.
At the other extreme, in the twentieth century a number of painters
have explicitly aimed at avoiding the impression of a perceived pictorial
space, a famous instance being Kasimir Malevich’s black square. Again
this phenomenon may in certain cases depend on viewing conditions. The
impression of a perceptual space may not be present, for instance, when a
painting is viewed from a very short distance.
Both extremes mark the poles of a continuum of perceptual experi-
ences. In between these poles we find the dual experience mentioned
before. In the present context, we will only speak of picture perception
proper, if both aspects are perceptually present to some degree.
83 The Dual Nature of Picture Perception

An Extension of the Standard Scheme


As pointed out before, the orthodox scheme of perception is inadequate
for this situation. An appropriate extension of that scheme is outlined in
figure 4.1. In this scheme we refer to the standard setting, where the phys-
ical stimulus consists of a flat pictorial surface.
It is worth noting, however, that the perceptual duality characteristic of
picture perception at the level of the percept is not necessarily tied to this
standard setting. That is, it is not restricted to situations in which the per-
ceived picture plane corresponds to a real surface while the perceived pic-
torial space is illusory. For such an impression could just as well be caused
by other stimuli. In contemporary art, for instance, James Turrell and
Amish Kapoor impressively demonstrated how an illusory percept of a
planar opaque object on a wall or the ground as part of perceived real
space can be evoked by a suitable arrangement of architecture and light-
ing. Because these illusory surfaces appear more or less homogenous in
color, no dual aspect is present in these cases. However, situations are
conceivable where an illusory perceived picture plane goes hand in hand
with a perceived pictorial space. In principle, even the entire perceived
“real space” could be illusory if a suitable (future) binocular virtual real-
ity device were used. At the other extreme, one could think of cases where
neither the perceived picture plane nor the perceived pictorial space are
illusory. Imagine, for example, an observer standing inside a camera
obscura looking at a translucent screen which is located in between the
observer and the camera’s pinhole (assuming some device which compen-
sates for the optical inversion). In this case the observer will have the
proper impression of a picture, with the perceived pictorial space corre-
sponding to the real scene in front of the camera obscura.
These final observations once again sustain our initial claim to the effect
that an appropriate account of both aspects of the dual nature of picture
perception necessitates referring to the percept itself. This feature of the
percept needs to be carefully distinguished from how—that is, by which
distal stimulus—it is generated. In consequence, from a psychological view-
point the emergence of dual percepts can in principle be discussed without
wedding it to the notoriously difficult concept of veridicality from the
start.

Related Forms of Perceptual Duality


What makes duality phenomena an important topic for general struc-
tural accounts of perception is the observation that they are not the least
84 Reinhard Niederée and Dieter Heyer

Figure 4.1
Extension of the standard framework of perception for the case of picture percep-
tion. As for the proximal aspect, see the section “Related Forms of Perceptual
Duality.”
85 The Dual Nature of Picture Perception

restricted to the case of picture perception. Obvious other examples are


mirror images and, in certain cases, perceived illumination patterns such
as shadows and slide projections. In the case of a mirror, typically a dual
percept is evoked in which the impression of a glossy, metallic, and opaque
surface, belonging to perceived real space, coexists with the impression
of a spatial configuration “behind the surface,” the so-called mirror
image. Similar phenomena occur with other kinds of glossy surfaces such
as water. Trompe l’oeil–like effects can occur if mirrors are looked at
obliquely, meaning that the mirror surface is not perceived, while the
mirror image is perceived as a part of real space.
In most situations, perceived shadows on a surface are merely an aspect
of perceived illumination. As mentioned before, this phenomenon alone
will not count as an instance of duality. Sometimes, however, the shadow
additionally takes on an objectlike or even spatial quality, which may then
be counted as an instance of duality in perception. The same phenomenon
can be encountered with highlights on a surface. A limiting case is a per-
ceived projected slide, where a perceived illumination pattern on the wall
may coexist with a full-fledged perceived pictorial space.
Even though the case of mirrors and objectlike shadows exhibits essen-
tially the same kind of perceptual duality found in the case of picture per-
ception proper, there are also some interesting differences concerning the
perceptual relation between perceived real space and the respective coun-
terparts of perceived pictorial space (e.g., a mirror image). For in these
cases, perceived real-world objects and certain perceived dual-world objects
often are linked perceptually in the sense that the latter are perceived as
being caused by the former.
Finally, perception in general comprises a kind of duality of aspects,
pertaining to the percept, which bears some resemblance to the duality
discussed before, even though in some respects it is fundamentally dif-
ferent in nature. These aspects are sometimes addressed as a “proximal
aspect” or “painter’s mode,” on the one hand, and the “distal” or “world
aspect/mode,” on the other (see, e.g., Rock 1983; Todorovic 2002).
Whereas the latter refers to the perceived scene (perceived-objects-in-
space), the former is meant to capture aspects of the percept that are
thought of as being more closely related to the proximal stimulus. Take,
for example, a situation in which an observer is looking at a disk that is
tilted relative to the direction of view. The observer will then perceive a
tilted circular object (distal aspect) and at the same time will have the
impression of a more or less elliptical contour (proximal aspect). The
86 Reinhard Niederée and Dieter Heyer

latter aspect closely resembles the perceived contour that a corresponding


picture drawn in perspective would display. Note that here the planar
aspect of the perceived picture is referred to. Nevertheless the proximal
aspect should not be identified with the planar aspect in picture per-
ception, because only in the case of a perceived picture do we also have
the impression of a planar surface in space. In consequence in a more
complete account of picture perception, three aspects could be distin-
guished: first, the proximal aspect and, second, the distal aspect which,
in turn, splits up into the planar and the spatial aspects. The necessity
of such a triple account becomes obvious when pictures are looked at
obliquely.

4.3 WHEN AND WHY DO DUAL PERCEPTS EMERGE? SOME


REFLECTIONS

Having outlined our concept of duality in picture perception and in other


perceptual phenomena, we now want to point out why we believe that per-
ceptual duality provides an important clue to our understanding of basic
features of perception in general. In particular, we will briefly discuss
some ideas about the functional role perceptual duality might play in
achieving perceptual stability in the presence of ambiguity. To do so, we
first need to take another and slightly more detailed glance at the general
scheme of perception discussed earlier.

The Concept and Problem of Inverse Projection


One of the milestones in the development of theories of vision was Kepler’s
theory of light rays projecting onto the retina, thereby forming the so-
called retinal image (cf. Lindberg 1976). By his account, the eye works
essentially in the same way as a lens camera. His theory made it possible
to integrate earlier theories of vision based on the concept of central pro-
jection into a modern optical framework. Before, the theory of central
projection had already played an essential role in Renaissance painting,
where it achieved an impressively mature state (which, in turn, was fore-
shadowed by ancient theory and art). Given a real or imagined scene, the
goal was to produce a painting that, ideally at least, yielded the same pro-
jection to the eye, and hence the same visual impression. In principle at
least, this method provides a powerful means to generate equivalent stim-
uli, that is, stimuli that generate identical retinal images and therefore
87 The Dual Nature of Picture Perception

have the same effect on the visual system (when they are viewed monocu-
larly from a specified viewing position and in certain fixed illumination
conditions). From a modern viewpoint, the concept of equivalent stimuli
refers solely to the physical principles connecting the distal and the prox-
imal stimulus. Although this level of analysis is sufficient if merely the
Renaissance painters’ goal is pursued, it falls short of providing us with a
perceptual theory that allows us to predict what is seen when we are look-
ing at a scene or a realistic painting and to explain why we see what we see.
The first problem that comes to mind is that, given a retinal image, there
is an indefinite number of—equivalent—possible scenes (seen from a fixed
viewing position) that could have generated this proximal stimulus. If per-
ception is understood as generating a scene-representation then the ques-
tion arises, which of these possible scenes should be represented? These
scenes might be called the possible virtual spaces associated with the prox-
imal stimulus. Of course, this question not only arises in the context of
picture perception but applies also to perception in general. Consider, for
instance, a cube viewed from a certain position. Usually, we will perceive
it as a cube. This cube could now be distorted in many different ways with-
out changing the retinal image. Hence, in all these cases we would have the
same percept; that is, we would perceive these objects as cubes. The ques-
tion now arises of why, given only that proximal stimulus, our visual sys-
tem generates a representation of a cube rather than a representation of
any of the other compatible objects.
If one would now produce an equivalent trompe l’oeil, we would of
course also see a cube. What the theory of central projection underlying
Renaissance art would then explain is that we perceive the same in both
situations, that is, when viewing the real cube and the picture of the cube,
but not why we see a cube at all. For when viewed from a fixed viewing
position, the picture (i.e., the proximal stimulus generated by it) is associ-
ated with an indefinite number of possible virtual spaces. Although it is
fairly unproblematic to speak of the perceived virtual (or pictorial) space
for a given observer at a given time, it seems unjustified to speak of the vir-
tual space of a picture independent of an observer. Unfortunately, many
investigators start out with the idea of a fixed virtual space of a picture,
directly or indirectly derived from the original scene underlying the pic-
ture (e.g., the scene photographed), which is then taken as a standard of
reference in their evaluation of the judgments made by their subjects con-
cerning their perceived virtual space.
88 Reinhard Niederée and Dieter Heyer

George Berkeley (1709/1975) and other early scholars in the field of


vision already clearly recognized the fundamental problem that the
just mentioned nonuniqueness throws up for any theory of vision. In
computational accounts of vision this difficulty is known as the ill-
posedness of the (unconstrained) “inverse projection problem.” As for
human vision, the situation seems to be even worse insofar as the actual
percept need not even correspond to any of the possible virtual scenes
associated with the retinal image (leaving aside the question of what
correspondence could mean here exactly). As a consequence, a theory
of vision has to face the major challenge of spelling out the principles
according to which our visual system constructs a more or less well-
specified—simplest? most likely?—perceived scene from the proximal
stimulus. This includes an analysis of why and how the visual system
organizes our perceived world into meaningful wholes, that is, perceived
objects, in the first place. For in the light pattern reaching the eye these
wholes are not present in any immediate way. So, strictly speaking, the
theory of central projection underlying Renaissance art is silent about this
construction aspect as well.

What Is Wrong with the Concept of Equivalent Stimuli?


On closer inspection, it turns out that the Renaissance explanation of why
realistic paintings work, based solely on the theory of central projection,
is in fact insufficient. For the strict notion of equivalent stimuli evoked in
this context is not applicable. This is because a painting and a depicted
scene hardly ever generate identical proximal stimuli. Ironically, this dis-
crepancy is most obvious in the case of line drawings based on contours,
which are commonly used to illustrate the geometry of projection under-
lying Renaissance art. To understand why we perceive such a drawing in
the way we do, we need to go beyond the physical notion of equivalence
with respect to the retinal images generated. Rather, a perceptual notion
of equivalence is needed in which the constitution of perceived objects on
the basis of their contours plays a central role. In other words, if one does
not want to rely merely on the experience that “it works,” which may be
sufficient for an artist, a theory of vision is needed to understand this
weaker notion of equivalence. However, even in seemingly less problem-
atic cases (photography or photorealistic paintings) the strict notion of
equivalence typically fails. For instance, the spectral composition and the
intensity of the light reaching the eye (from the picture and the depicted
89 The Dual Nature of Picture Perception

scene, respectively) in most cases are clearly different, with contrast being
severely reduced in the former case.
Even if it were possible to achieve strict equivalence, the usefulness of
this concept is limited from the start because it is restricted to monocular
viewing and does not take into account the dynamics of perception. That
is, the observer’s viewpoint has to be fixed, thereby excluding bodily
movements, which are so important in ordinary vision. Even in the case of
monocular vision with a fixed viewing position, accommodation intro-
duces a dynamic element, yielding different results for an observed paint-
ing and the corresponding scene.
Last but not least, the fact that most realistic paintings evoke a dual
percept clearly shows the nonequivalence of the picture and the corre-
sponding scene. For, if they were equivalent, a trompe l’oeil effect would
have to occur. In other words, one would have the visual impression of a
real scene as viewed through a window. Instead, the visual system comes
up with a percept in which two perceived virtual spaces—one of them
being the picture plane in a certain position in space—are intertwined, as
it were. Why does this happen? Rather than being a nuisance, the afore-
mentioned difficulties turn out to be a clue for our theoretical under-
standing of the dual nature of picture perception and possibly even of
some basic structural features of perception in general.

Cue Integration and Cue Segregation


Before turning to the question just posed, we will first briefly sketch some
features of the standard account of depth perception in general and how
it is usually applied to perceived depth in pictures. On this account, the
perception of spatial features in a scene is conceived as being based on a
number of so-called depth cues. Some of them already apply to static
monocular viewing, such as shading and cues derived from the geom-
etry of central projection. The latter include occlusion, texture gradi-
ents, height in the projection plane, and so-called linear perspective cues
(referring to certain linear image elements, such as converging lines inter-
preted as parallels in space). Other monocular cues involve dynamic
components, such as the cue of motion parallax and the ocular-motor cue
of accommodation. Finally, we have cues that only work for binocular
vision, such as binocular disparity and the ocular-motor cue of conver-
gence. For a more detailed review see, for example, Stephen Palmer
(1999).
90 Reinhard Niederée and Dieter Heyer

Considered individually, each of these cues, if applicable at all, tends to


yield depth information of only limited reliability and strength. This sug-
gests the assumption, shared by most researchers, that the visual system in
one way or another combines various depth cues to create a sufficiently
reliable and sufficiently specific scene representation. In the recent com-
putational literature on spatial features of perception this idea has led to
various formal models of “cue integration” (see, e.g., Landy et al. 1995,
who also review some of these models and related Bayesian approaches).
The idea of cue combination is commonly employed as an explanation
for what will be called here the monocular depth enhancement in picture
perception. This is the well-known observation that in most cases the per-
ceived depth pertaining to perceived pictorial space is markedly more dis-
tinctive, or pronounced, when pictures are viewed monocularly, from a
fixed viewpoint, and through a reduction tube that shields off the picture
frame. (In fact, in many cases a similar phenomenon is found for colors,
illumination, and gloss in perceived pictorial space.) Only recently has
monocular depth enhancement been studied quantitatively in a series of
studies (Koenderink, van Doorn, and Kappers 1995; Koenderink et al.
2001).
The standard explanation for monocular depth enhancement is based
on the assumption that under those constrained viewing conditions, con-
flicting depth cues that under ordinary viewing conditions signal flatness
(e.g., binocular disparity, motion parallax) are eliminated. In ordinary
viewing, in contrast, these conflicting cues are combined with the static
monocular cues responsible for the impression of pictorial space, yielding
a single compromise scene representation.
In being confined to the concept of a single scene representation as the
outcome of a cue combination process, traditional cue integration theories
are by their very nature unable to account for duality phenomena in per-
ception. However, what needs to be accounted for in picture perception
is not only the monocular depth enhancement phenomenon but, first and
foremost, the duality of picture perception. One could of course contem-
plate extending the standard account accordingly. Let us roughly sketch
one general way of how this could be done. Assume that given a cer-
tain proximal stimulus, the visual system encounters a situation in which
considerable conflicts between cues occur, and two (or possibly more)
sufficiently rich coherent subsets can be found, each of which is apt to trig-
ger a scene representation of its own. Then in certain cases the visual
system will constitute corresponding subpercepts that coexist, rather than
91 The Dual Nature of Picture Perception

just form a kind of simple compromise or decide on only one of them. The
formation of different subpercepts, each of which involves a process of cue
integration, goes hand in hand with a kind of cue clustering we will call
cue segregation.
These different subpercepts are not simply generated in isolation but
there are mutual interactions at various levels of processing. Adopting a
term coined by Sedgwick (chapter 3 this volume), we assume a cross-talk
between those processes. For instance, the reduction of the distinctive-
ness of perceived depth in ordinary picture viewing (as opposed to static
monocular viewing) might be explained by a cross-talk between (pro-
cesses generating) the perceived picture plane and the perceived pictorial
space. The same goes for the often reported inverse phenomenon that in
certain situations the planar aspect of a perceived picture seems less vivid
in the presence of a perceived pictorial space that involves a strong impres-
sion of depth. (Further potential instances of cross-talk are described later
on and in chapter 3 in this volume.)
Another kind of interlock between the generated subpercepts is some
kind of phenomenal binding, which in the case of picture perception is
experienced as the perceptual unity of a perceived picture. The perceived
pictorial space and the perceived picture plane perceptually belong to
each other, as it were. For certain types of duality, the visual system even
seems to create a suitable type of gestalt (e.g., “perceived picture”), which
due to learning possibly becomes equipped with specific higher-order cues.
Furthermore, as already mentioned before, such subpercepts may be
experienced as standing in a hierarchical order to each other, only one of
them belonging to perceived real space.
Like any other account of perception, this scheme needs of course to be
complemented by the concept of attention. In particular, attention may be
shifted between different subpercepts, possibly leading to a refinement or
enhanced vividness of one subpercept at the expense of another.
Finally, one might contemplate integrating the concept of a proximal
mode into our scheme by assuming that it is primarily related to an initial
state of processing that forms the basis for the later extraction of cues
and precedes the processing of cue integration and segregation. The pre-
vious remarks on cross-talk, phenomenal binding, and shifts of attention
mutatis mutandis apply here as well.
Needless to say, our proposal of how duality concepts might be inte-
grated into common concepts of cue integration throws up more ques-
tions than it is able to answer. Nonetheless, we are convinced that duality
92 Reinhard Niederée and Dieter Heyer

phenomena in perception should not simply be neglected and therefore


that they should be integrated into concepts of cue integration in one way
or another. Maybe, the cue concept will itself eventually turn out to be
inadequate, after all, which then would call for a still more fundamental
revision of the general conceptual framework under consideration.

Duality in Perception: Functional Aspects


At first sight, duality in perception might look like a defect—at least, when
the visual system is viewed from a perspective that emphasizes its repre-
sentational features. Why should the system come up with two or more
different interpretations of the proximal stimulus at the same time? We
suggest that duality is one of a variety of strategies our visual systems
employs to cope with perceptual ambiguities. In one way or another, these
strategies allow the visual system to represent (at the level of the precept)
the possibility of certain alternative representations compatible with the
proximal stimulus. On the standard account of perception in contrast,
the visual system would always have to decide for one specific scene-
representation, regardless of what kind of input it gets (even though
perhaps nobody actually will maintain such a strict position). When
conflicting or insufficient cues are present, this could in certain situations
lead to unreliable scene-representations (illusions) or to the oscillation
between different scene-representations (bi- or multistability). In fact, our
visual system appears to reduce the frequency of risky illusions or confus-
ing multistability by a number of strategies, such as the following ones
(cf. figure 4.2):
1. Perceived occlusion. Here, we have an overall impression of the par-
tially occluded object (so-called amodal completion; cf. Michotte, Thinès,
and Crabbé 1964/1991). At the same time we have the visual impression
that the occluded parts are not visible, which leaves open other possibili-
ties of what could be behind the occluding object. Note that we are not
simply speaking here about the physical fact of one object being occluded
by another relative to the eye but of perceived occlusion.
2. Perceptual vagueness. Think, for instance, of an object seen through
fog or of the diffuse character of the perceived size and distance of objects
far away.
3. Impossible figures. Here a visual overall impression emerges that lo-
cally conforms to physical scene descriptions but which is globally incon-
sistent as a global scene description.
93 The Dual Nature of Picture Perception

Figure 4.2
Strategies of how the visual system copes with perceptual ambiguities. The two
extremes (top/bottom) correspond to two stimuli that evoke different clear-cut
percepts. As an example, the case of picture perception is considered. If, now, one
proximal stimulus is gradually transformed into the other, conflicting cues will
arise. Some of the responses of our visual system are listed on the right.

4. Interplay between proximal and distal mode. Think, for example, of a


house seen through a window, which at the level of the percept simulta-
neously appears smaller than the window (proximal mode) and larger
than the window (distal mode).
5. Duality. Various forms of perceptual duality, such as those discussed
earlier (pictures, mirrors, objectlike shadows).
Of course, several of these phenomena, including bistability, may occur
in combination. Simple examples are pictures of an object in fog (duality,
vagueness) or the picture of a Necker cube (duality, bistability).
94 Reinhard Niederée and Dieter Heyer

These examples suggest that our visual system possesses a rather sophis-
ticated ambiguity management up to the level of the percept, which so far
has hardly been taken into account in a systematic way.

4.4 FURTHER SHORTCOMINGS OF THE WINDOW METAPHOR IN


PICTURE PERCEPTION

Let us focus attention again on picture perception proper. As pointed out


before, the classical window analogy is at least incomplete, insofar as it
fails to accommodate the dual nature of picture perception. One might
now contemplate maintaining the window metaphor by applying it only
to the spatial aspect of perceived pictures. This would then imply that the
planar and the spatial aspect correspond to different inverse projections.
We will now discuss two phenomena that show that even this more lib-
eral interpretation of the window analogy is inappropriate. Interestingly,
these phenomena seem to be restricted to the case of picture perception
proper, that is, they are not encountered in trompe l’oeil situations. In
consequence, these intricacies seem, indeed, to be intrinsic features of the
dual processing of pictures.

Compensation Hypotheses versus Cross-Talk


First, we will deal with the question of what happens when pictures are
viewed from different positions. For each viewing position, a different reti-
nal image results, each of these associated with a specific set of possible
virtual spaces in terms of central projection. According to the window
analogy, corresponding changes of the perceived pictorial space would
have to occur when the viewing position is changed. There is some evi-
dence that the observed changes in many situations appear to be less than
expected, however. We will call this the robustness of perceived pictorial
space (for a review of some of the relevant literature see, e.g., Rogers
1995).1 According to Pirenne (1970), changes that appear to be in line with
the window metaphor occur in the case of trompe l’oeil (e.g., with Pozzo’s
aforementioned ceiling in Sant’Ignazio). When viewed from a viewpoint
other than the station point, drastic distortions may be perceived, which
are much stronger than those ordinarily experienced in picture perception
proper, that is, when a dual percept is evoked (where the station point is
an ideal viewpoint corresponding to the center of projection).
A number of researchers have proposed an explanation of such phe-
nomena in terms of a so-called compensation mechanism (such as Gibson
95 The Dual Nature of Picture Perception

1947/1982, Pirenne 1970, or Kubovy 1986; for a discussion, see Rogers


1995). In a nutshell, such compensation hypotheses assume several stages
of visual processing assumed to work roughly as follows. First, the picture
surface and how it is situated in space are determined; building on that, a
“correct” viewing position is determined by the postulated mechanism,
relative to which the visual system finally constructs the pictorial space.
Such approaches appear problematic for various reasons. The theoretical
question arises of what the “correct” viewing position should be and
according to what principles the visual system should construct such a
point. Also, the empirical adequacy of these approaches seems question-
able. Ideally, by such an account the visual system would always con-
struct the same viewing position relative to the perceived picture plane
and hence the same associated pictorial space. This is obviously not the
case. An interesting extreme case is that of anamorphic pictures. Consider,
for example, an anamorphotic painting of a skull as in Hans Holbein’s
famous Ambassadors (1533). From the correct station point a skull is
seen, whereas from a frontoparallel viewing position we have the impres-
sion of a pretty long egglike object. This is clearly at odds with the com-
pensation prediction.2 Finally, it does not seem plausible that the visual
system is endowed with a complex mechanism just for the case of picture
perception.
Instead of assuming such a compensation mechanism, we hypothesize
that the aforementioned phenomena may to some extent be due to a cross-
talk between the processes yielding perceived pictorial space and the per-
ceived picture plane, respectively. In cases where the perceived picture
plane—thanks to shape constancy mechanisms—appears more or less
invariant, the robustness of pictorial space found in many situations might
be due to an influence the latter exerts on the former (for a similar posi-
tion, see chapter 3 in this volume).
Note that this concept of cross-talk (which without doubt is in need
of a much more detailed elaboration) has the interesting property that it
could in principle work both ways. Consider, for instance, the example of
the anamorphic skull, which on the picture plane corresponds to a long
egglike patch. Recall that from the station point, the skull is seen. At
the same time it is virtually impossible to see a long egglike patch on the
picture plane; instead, a patch whose shape resembles the contour of
the skull is seen. This might be due to a reverse cross-talk effect, namely,
an influence perceived pictorial space exerts on the perceived picture
plane.
96 Reinhard Niederée and Dieter Heyer

Finally, recall that compensation theories usually assume that perceived


pictorial space corresponds to some scene representation (viz. one of the
possible virtual spaces associated with a “corrected” viewing position). In
contrast, we assume that due to cross-talk and other factors, “incoherent”
perceived pictorial spaces may emerge, that is, spaces that do not corre-
spond to any possible scene representation. This incoherence seems to be
one of the reasons pictures look “strange” from certain viewing posi-
tions. Certain experimental findings reported in the literature (see, e.g.,
Rogers 1995) may be interpreted this way. Last but not least, such intri-
cacies highlight the fundamental conceptual and methodological difficul-
ties notions of correspondence between the physical world and perceptual
experience raise from the start.

Dissociation of Spatial Reference Systems


We now turn to the second feature of picture perception that is not in
line with the more liberal reading of the window metaphor. In everyday
situations our visual system usually establishes, roughly speaking, an
environment-centered spatial frame of reference. This includes the basic
orientations of the horizontal and the vertical. In addition we have a per-
ceptual awareness of ourselves as looking at objects from a certain direc-
tion. We may, for instance, have the impression of looking down to the
floor, which itself appears horizontal. The same thing happens when we
are looking through a window. Typically, we then have a unifying refer-
ence system that connects the perceived world in front of and behind the
window.
The same kind of integration is achieved by our visual system in the case
of a trompe l’oeil situation. Imagine, for instance, that we are looking at
a trompe l’oeil from the station point. Imagine further that the trompe
l’oeil is painted in such a way that in our perception the room we are
standing in (along with its floor) continues in perceived pictorial space
(i.e., ideally there is no perceptual distinction between perceived real and
perceived pictorial space; an example of such a trompe l’oeil is the fresco
by Baldassare Peruzzi in the Salla della prospettiva, Villa Farnesina, Rome
[see Kubovy 1986]). Assume now that the trompe l’oeil was moved up the
wall in such a way that its bottom now is above our head. If a trompe l’oeil
effect would then still occur, what we would see then—in line with inverse
projection—would be a ground plane tilted down toward us. The visual
system would maintain a common frame of reference: the perceived
ground plane belonging to pictorial space would no longer appear hori-
zontal, and we would have the visual impression of looking at that plane.
97 The Dual Nature of Picture Perception

Consider now a similar situation, where we look at a picture, but no


trompe l’oeil effect occurs. When the painting is now moved up, we typi-
cally still have the impression of a horizontal ground plane (in conflict
with inverse projection) which appears to us as being seen from above
while at the same time the picture plane is perceived as being seen from
below. In other words, we here get a dissociation between the spatial frame
of reference pertaining to perceived real space (including the picture
plane), on the one hand, and perceived pictorial space, on the other.
Metaphorically speaking, it is as if in picture perception proper our
visual system takes a separate virtual viewpoint associated with perceived
pictorial space (analogous to the “mental eye” in Koenderink et al. 2001).
As demonstrated by our everyday experience of looking at a photograph
in a newspaper lying horizontally on the table in front of us and many
other situations, this phenomenon extends to a variety of possible orien-
tations and positions of the picture plane. Similar remarks apply to per-
ceived size and distance in pictures. The dissociation observed in these
cases goes hand in hand with a relative stability of the spatial frames of
reference associated with perceived pictorial space, which in turn is an
important aspect of the robustness of pictorial space.

4.5 EPILOGUE

The preceding analysis gives only a first impression of the richness of the
perceptual phenomena involved in the perception of pictures and, to some
extent, mirrors and shadows. Further complexities arise when pictures are
painted or projected on curved or transparent surfaces. In those situations
sometimes a complex phenomenal interlock between perceived pictorial
space and perceived picture surface occurs, which no longer can simply be
described as the coexistence of two neatly separable aspects (several
examples of this can be found in contemporary art, e.g., in the work of
Tony Oursler). Another type of complexity arises when pictures involve
multiple perspectives, or when their rendering becomes less realistic, as for
example in impressionist art. Last but not least, the dual-ity of the spatial
and the planar aspect extends to the dynamic case of motion pictures.
Without doubt, a much more refined theoretical account, thorough phe-
nomenological studies, and experimental investigations are needed to
obtain a deeper understanding of such phenomena in picture perception.
Once one has an increased awareness of duality in perception, a
plethora of analogous phenomena that go beyond the dichotomy of a
spatial and a planar aspect attracts one’s attention. Suffice it to briefly
98 Reinhard Niederée and Dieter Heyer

mention a few examples. First, duality can occur at various levels simul-
taneously. Think, for example, of perceived pictures in a picture or paint-
ings like Givseppe Arcimboldo’s Summer. Here one perceives (as a
part of perceived pictorial space) a head composed of fruit and vegetables.
That is, at the same time a head and an arrangement of fruit and vegeta-
bles are seen. Or think of a wooden sculpture of a woman, where, loosely
speaking, the impression of a woman and of a piece of wood coexist. In
theater the simultaneous perception of an actor on the stage and of a
character within the drama “put on stage” may be counted as an instance
of dual perception (e.g., Polanyi 1970b). Of course, in all these cases, an
additional symbolic aspect can come into play that increases complex-
ity even further. All in all, perception seems to embody some of the com-
plexities typically thought of as being characteristic of language-based
cognition. In certain respects, at least, our perceptual experience some-
times seems to resemble a multilayered narration more closely than a
straightforward representation of “what there is.”

Notes
1. Note that in order to verify “that a change is less than expected,” one would,
strictly speaking, have to compare the perceived pictorial space with all of the pos-
sible virtual spaces associated with the retinal image. As a measure of discrepancy
one would then have to take the distance to the best-fitting virtual space, as it were.
Unfortunately, most studies in the field base this measure only on a single distin-
guished virtual space associated with the respective viewing position. To this end,
a “true” virtual space is singled out associated with a standard viewing position
(station point). This true virtual space is then geometrically transformed into
“the” virtual spaces for other viewing positions. So far we have not found an ade-
quate justification for this approach.
2. If the visual system were able to reconstruct the correct viewpoint, a skull
would be seen in both cases. If, in contrast, the system always picked a virtual
viewing position perpendicular to the perceived picture plane, the long egglike
object would be seen in both situations.
Chapter 5
Perceptual Strategies and Pictorial Content

Mark Rollins

5.1 INTRODUCTION

Recently in cognitive science there has been a healthy trend toward rethink-
ing some old debates. In particular, the traditional opposition between
constructivism and direct realism—the view that perception depends on
mental representations versus the view that it does not—has been cast
aside by several philosophers and scientists in favor of what can be seen
as a hybrid account. Such hybrids go by a variety of names: “animate”
or “utilitarian” or “interactive” theories of vision, for instance. (Ballard
1991; Ramachandran 1990). But all are distinguished by their rejection
of central ideas in the work of David Marr. Although Marr’s “computa-
tional Gibsonism,” as he called it, is already a compromise, it combines
elements of constructivism and direct realism in the wrong way: it restricts
the range of knowledge on which the visual system can draw and builds
certain assumptions into it. Such constraints have the effect of making
the perceiver respond to visual information partly as if the response were
direct, because no variation is permitted across perceivers. There can be
no difference due to background knowledge; thus the response is to some
extent determined by causal law. But constraints of that sort merely pro-
vide a framework within which elaborate representations are constructed.
What is wrong with Marr’s compromise, according to recent research, is
that it presupposes that such rich representations are needed; and some
evidence strongly suggests that they are not.
An alternative that has recently emerged makes vision depend on rep-
resentations but of a rather sketchy sort: “visual semiworlds,” as they have
been called (Churchland, Ramachandran, and Sejnowski 1994). These
are sufficient, it is said, precisely because perception is not constrained in
Marr’s way. The visual system draws upon background knowledge of
100 Mark Rollins

various kinds, and it collaborates with other sensory systems and motor
control: the eyes conspire with the ears and with muscles, tendons, and
joints to enable the viewer to take representational shortcuts. The virtue
of such a cooperative sensory scheme is that it makes possible perceptual
strategies. With those in hand, precise computations and representations
are no longer required.1 I shall refer to a theory of this type as a strategic
design theory (SDT). The primary question I want to address is how SDT
might be applied to picture perception and, in particular, to the perception
of pictorial space in the context of visual art.

An Example: Motion Capture


It is well known that a series of flashing lights can produce the appearance
of motion. But, as Vilayanur Ramachandran (1990) points out, it is also
the case that even if one light is occluded, it will appear to move behind
the occluder if it is located properly in the context of other alternately
flashing lights. This is an instance of a class of cases that we can call
(following Ramachandran) capture phenomena: the location of the miss-
ing light and its motion from one point to another are captured, as it were,
by the other lights, although the absent stimuli cannot be actually regis-
tered in perception. Importantly, this effect cuts across different sense
modalities: alternating the location of sounds, one of which corresponds
to the position of a flashing light, can also produce the appearance of
motion to the point behind the occluder. Here the visual information is
captured by auditory stimuli.
Motion involves a change of position, of course, and a spatial trajec-
tory from one place to the next. Where the moving stimulus is complex,
there is also the problem of defining a spatial register between features of
the stimulus at one moment in time and the same features a few moments
later. This “correspondence problem” is one, Ramachandran argues,
which motion capture is able to sidestep. To cite his central example, if
this account is correct, it will not be necessary for the visual system to
compute the trajectories of each and every spot on a leopard as it leaps
from tree to tree.
It is clear that knowledge plays an important role in these cases;
namely, in the assumptions that are needed to disregard information. For
instance, as Ramachandran notes, the utilitarian solution to the corre-
spondence problem in motion detection presupposes the knowledge that
spots do not normally jump off of surfaces or otherwise move around
discontinuously. Such knowledge is part of a conception of the physical
101 Perceptual Strategies and Pictorial Content

world, having to do with properties of surfaces. Nonetheless, there is an


important sense in which such assumptions or constraints are ecologically
specific. Although the constraints may be inherent in either the nature of
the task or the visual equipment available to the perceiver, different con-
straints can be brought to bear by the different strategies employed to per-
form the task. And the training of the perceiver’s abilities in this regard
will be context dependent. Thus, which tasks are governed by which con-
straints can vary from perceiver to perceiver.
In any case, what strategic design theories explicitly deny is that knowl-
edge of this sort can be used to compute or infer exact locations of each
and every spot. It may be that the relevant knowledge in this case is
really procedural; it involves knowing how to move a cluster of points or
regions in space along a trajectory, keeping them together without count-
ing them, in effect. If so, the knowledge need not be encoded (certainly not
symbolically) but only have the status of a conditioned (internal) disposi-
tion or ability. But even if the knowledge is theoretical, vested in declara-
tive or semantic memory, the point is that it does not specify, nor is it used
to compute, all of the possibly relevant details. At any rate, the claim of
the hybrid approach is not that no representation is required for percep-
tion at all but only that fewer or less complete ones will do. The point is
that a perceptual task is performed without a detailed representation of
the sort on which Marr-style computation would depend.
What, then, does motion capture have to do with static picture percep-
tion of the sort required for visual art? The answer is that in the SDT
approach, nature generally favors an economical use of resources. This is
apparent in the case of motion capture as a way of handling the corre-
spondence problem, and it will emerge that the same principle in a differ-
ent form is supposed to be at work when we process the information in a
painting or drawing. But before we turn to that, I first want to consider
the sense in which SDT might be said to view perception as partly direct.

5.2 WHAT IS A HYBRID PERCEPTUAL THEORY?

The claim that picture perception is partly direct means different things to
different people. It is sometimes said, for instance, that picture perception,
even from a Gibsonian perspective, is indirect, although the processes by
which the information in the picture are accessed involve no elaborate
internal representations (e.g., Sedgwick, chapter 3 in this volume). But
this means only that perception of the depicted scene is mediated by the
102 Mark Rollins

picture, which can itself provide information directly. In the words of a


philosopher whose views are congenial to directness: “Pictures are visual
prostheses; they extend the informational system . . . in ways that depend
upon and also augment our ability to identify things by their appearance”
(Lopes 1996, p. 144). According to this view, pictures are external records
of events that make internal records, at least in the form of detailed copies,
unnecessary. In this case, when it is said that “picture perception” is indi-
rect, that simply means that objects, scenes, or events are perceived by way
of a picture. We can therefore call this the pictorial surrogate view.
However, sometimes it is said that picture perception is indirect (accord-
ing to James J. Gibson), and more is meant than that pictures mediate the
pickup of information, which the depicted scene would also contain (e.g.,
Cutting and Massirioni 1998). The claim, in this case, means that access-
ing that information in the picture may, in fact, require internal represen-
tations. However, based on this view, whether or not that is so will depend
on the nature of the picture. Sometimes picture perception will require
internal representations, and sometimes it will not. Thus the “picture per-
ception” that is partly direct, according to this use, is really the generic
ability and not any particular instance of it; and “partly” means “for some
pictures, but not all.” Call this the occasional constructivist account.
Neither of these views entails a hybrid theory of the sort that I wish
to elucidate. Such a theory claims that individual pictures are perceived
through a mixture of direct and indirect processes. This view might be
taken to apply to all pictures. But it might also be more limited in scope.
For instance, we could adapt the occasional constructivist insight—that
pictures need not all be perceived using the same processes—but maintain
that pictures fall along a continuum, rather than being of two distinct
types. Purely direct access and wholly indirect perception would be found
at the ends of the continuum, with mixtures of various sorts in between.
This would amount to an occasional hybrid theory.2 My focus would then
be on the pictures in the middle of the range. Even there, however, there
is the need to investigate the sense in which the direct component of per-
ception might be direct.
At first glance, the sense in which SDT treats perception as partly direct
seems clear. In positing various ways in which the visual system econo-
mizes on resource use, SDT cuts back, in effect, on mental representations.
It envisions, for example, replacing the precise delineation of details with
spatial tracking of coarsely coded features to resolve the problem of
correspondence. It is possible, then, that the “direct component” of SDT
103 Perceptual Strategies and Pictorial Content

is no more than the absence of an elaborate mental representation and


the reliance on partial information contained in the momentary glance.
Perception would thus be direct just insofar as it is not indirect.3
However, even partial representations are internal media that stand
between the perceiver and the world, and they can involve a constructive
process, even if it only results in relatively simple forms. In any case, the
emphasis in SDT on active processes, such as the movements of eyes,
head, or body, makes it clear that perception is imagined to be partly
direct in a much more positive sense than the absence of elaborate repre-
sentations would suggest. Directness is identified, perhaps, with the mech-
anisms by which momentary glances are coordinated (or with their effects
on perception), rather than the mere fact that the information they con-
tain is fragmentary and incomplete. This suggests that we can construe the
directness of perception according to SDT in the following way: internal
representations are required for picture perception, but the roles they play
in perception are both inferential and noninferential in form. In the discus-
sion that follows, I will explain noninferential operations on representa-
tions in terms of perceptual strategies. It is by virtue of these strategic
operations that perception is direct. The fact that they can be defined
in terms of uses of representations makes SDT a hybrid theory in a very
strong sense.
Armed with that interpretation of SDT, then, we should be able to
define the issue of hybridity in a clear and precise way. There are, it
would seem, two opposing positions. On one hand, there is David Marr’s
account, which is often linked to Gibson’s by virtue of Marr’s emphasis
on constraints on computation. In the words of Michael Kubovy and
William Epstein (in press): “According to the approach of Marr and
Gibson constraints are neither (1) lodged in the mind nor (2) are they
active constituents in the perceptual process. They are the conditions
which the world must satisfy if the computational algorithms are to go
through.” The implication is that perception is direct to the extent that it
relies on bottom-up processes, constrained in this way.4 If pictures do,
indeed, come in a variety of types, we might then argue that they vary
according to how much weight they require to be assigned to bottom-
up, modularized processes, on the one hand, and to top-down, cognitive
processes, on the other. For some pictures (those midway along a con-
tinuum), both would be given significant weight.
On the other side, there is Ramachandran’s SDT-type account, which
rejects Marr’s stage-based approach and defines a composite theory in
104 Mark Rollins

terms of an inferential/noninferential process blend. The noninferential


ingredient, on this account, is what allows us to employ internal represen-
tations that are only partial and incomplete.
Unfortunately, the matter is not as simple as that dichotomy suggests.
There are complications on both sides of the fence. For one thing, even
proponents of a paradigmatic form of SDT, such as Ramachandran,
embrace modularity in a certain sense, at the same time as they are reject-
ing Marr’s modularized computational model. The operations of mod-
ules depend on assumptions and constraints. This means that the visual
system’s instantiation of such constraints does not preclude its reliance
on heuristics and strategies. If perception is direct to the extent that it
employs processes that are bottom-up, SDT suggests that it can also be
direct in another sense, namely, in the noninferential control processes by
which its resources are deployed. For SDT in the flesh, the two forms of
directness can hand in hand.
On the other side of the fence, there is the theory of James Cutting,
which he calls “Directed Vision” (1998). According to that account, infor-
mation in a scene is actually overspecified: there will be multiple sources
from which a perceiver can pick the information up. Thus the relation
between information and the properties by which it is specified is not,
as Gibson thought, one to one. Nonetheless, the “bottom-up” sense of
directness remains in play: whatever the source happens to be, it may pro-
vide the information without intervention from cognitive processes. Thus
Cutting describes direct specification in a way that suggests a close align-
ment with Marr: “[I]n many cases the specification of a percept by infor-
mation can be written as a deductive syllogism. . . . That is, given certain
assumptions (axioms), a definitive conclusion can be drawn” (1998, p. 86).
Further, “natural law is followed [in perception] because there is no alter-
native, and mathematics is used because that is the way perceptual sys-
tems work. The natural law and mathematics are deeply ingrained in the
nervous system, and their rules constrain and guide perception and cogni-
tion” (p. 75). As I have noted, Cutting maintains that the perception of
some pictures depends on inference and higher-order cognition, and this
opens the door to a Marr-style hybrid account, at least on occasion. At the
same time, the many-to-one relation that Cutting describes also implies
that perceivers can respond (directly) to the same information in different
ways: he speaks of different “styles” by which perceivers may select or
integrate information. Although he does not put it this way, here again
the two forms of hybridization appear to work together.
105 Perceptual Strategies and Pictorial Content

That fact is important for an understanding of pictures. In recent years,


it has often been argued that what a picture represents depends on what
can be recognized in it, using ordinary perceptual abilities (Schier 1986).
That is, no special training in the use of pictorial conventions is required.5
This Recognition theory, as it is called, implies that pictures are individu-
ated by the particulars of the psychological response they tend to gener-
ate. What any picture represents, its content in one sense, is determined by
what can be seen in it; and what can be seen in it depends on perceptual
processes. This is true, presumably, of both the type of object or scene that
the picture represents and the particular instance or token of that type: a
babe-in-arms, for example, as well as the infant Jesus held by its mother,
Mary. Insofar as picture perception is indirect, we are inclined to say that
the content of the picture is determined by the content of the internal rep-
resentations produced in response to it. There is, then, the difficult ques-
tion of how to individuate those internal representations. But Recognition
theory has been endorsed by some psychologists who are friendly to the
idea that perception is partly direct (e.g., Cutting and Massironi 1998).
If they are right, it must be possible to say how, on a recognition-based
account, the content of the picture is partly identified in terms of direct
information pickup. Recognition accounts of pictorial content draw
on Marr’s type of computational Gibsonism. But, as the complexities
described make clear, that leaves open the possibility that the directness
of picture perception might be understood in more than one way. My
interest is not really in the proper use of the term “perceptual directness.”
My aim is, rather, to indicate the impact on the theory of depiction that
exploring the non-Marrian construal of directness might have. What I will
argue is that the contributions of the two psychological components,
direct and indirect, to pictorial content are not what Recognition theories
have taken them to be.

5.3 PERCEPTION AND CONVENTION IN PICTURE RECOGNITION

I begin by recalling Ernst Gombrich’s famous claim that “there is no inno-


cent eye.” In saying that, Gombrich implies that perception is a matter of
convention, in the sense that perceptual categories vary from place to
place and that diverse techniques for representing them are taught and
learned. Recognition theories have also been used to link perception to
convention; but only by inverting Gombrich’s idea. They say that con-
vention depends on perception. That inversion is motivated by the goal of
106 Mark Rollins

reconciling a perception-based theory of depiction with Nelson Good-


man’s idea that art is a language. The claim means more than just that pic-
torial conventions are learned and applied by means of vision.
Goodman’s conventions are formal, of course, in a way that Gombrich’s
are not. Goodman claims that visual art represents by way of a system of
arbitrary symbols that have distinctive syntactic and semantic properties,
just as a verbal language does. The knowledge required to decode a pic-
ture is thus special and not derived from a general experience with the per-
ceptual world. Indeed, on the conventionalist view, we often learn how to
perceive the natural world from looking at pictures, rather than the other
way around. Pictures define space for us, they do not simply represent it.
In Goodman’s words: “Nature is a product of art and discourse” (1968, p.
297). However, according to certain recent accounts, the semantic prop-
erties of even a verbal language, at least some of them, must be explained
in terms of perceptual mechanisms that serve to locate the speaker’s body
in space. And it is the semantic properties of this sort that provide the
model for pictorial representation. In light of that, space cannot simply be
a pictorial convention, for the relevant conventions presuppose it. Visual
art might be a language, but the very properties that make it so must be
explained in terms of perceptual psychology.
The idea that convention depends on perception is most apparent in the
case of demonstrative reference (Evans 1982; Schier 1986; Lopes 1996).
When I refer to “this cup” or “the lamp over there,” I must establish per-
ceptual contact with the relevant object, both for myself as the speaker
and for listeners with whom I wish to communicate. This requires me to
look, point, or nod in a certain direction, thus linking the location of an
object to the position of my own body. When a listener looks in that direc-
tion, he, too, understands the reference by locating the object and me in
relation to himself.
In the case of pictures, even if reference is not exactly of the demon-
strative sort, the need for perceptual contact with the depicted object
by way of a spatial location is the same. According to Gareth Evans,
for example, pictorial reference is mediated by thoughts of a type that
depends essentially on early visual processes. In this he is followed by
Flint Schier, Christoper Peacocke, and Dominic Lopes.6 Thus Peacocke
describes a role for spatial “scenarios,” as conditions on perception that
are essentially ways in which a perceived scene is organized in relation to
the perceiver’s body: what is in front of or behind the perceiver, to the
perceiver’s right or left. Lopes, for his part, emphasizes spatial “aspects,”
107 Perceptual Strategies and Pictorial Content

which are “patterns of visual salience,” defined by the perceiver’s point of


view. Not all aspects are purely spatial, but even those that aren’t presup-
pose and depend upon the spatial organization of the picture. Pictorial
reference, on his and Schier’s account, is actually established by the causal
history of the picture.7 But the perceiver’s understanding of reference
is mediated by the totality of features depicted, which are themselves
dependent on the spatial aspect that the picture contains. To that extent,
pictures function, not so much like names for, or descriptions of, the
things they denote, as like verbal pointings to “this” or “that” object,
located over “here” or “there.” The basic idea is that pictures embody a
perspective or point of view that allows some features of the depicted
object to be seen and preclude the representation of others.8
The central assumption, shared by Peacocke, Evans, Lopes, and Marr,
is that the perception of basic spatial properties depends on internal
processes that are nonconceptual and unaffected by background knowl-
edge and beliefs. There is of course no general consensus about what
the term “concept” and cognates like “conceptual” or “nonconceptual”
mean. But the basic idea at work in these accounts is that concepts are
units of representation that play roles in the construction of propositions
and the drawing of inferences that depend upon domain-general knowl-
edge. As such, concepts are interdefined in a holistic framework, stored in
semantic memory. But none of that is true of spatial aspect representa-
tion. Conceptualization is eventually necessary to establish reference, but
that must be “grounded” on nonconceptual perceptual contact via the
spatial organization of the picture (Lopes 1996, p. 102, n. 12). That orga-
nization is encoded in elements that are unchanged by the higher-order
processes to which they contribute; perceptual content is thus atomistic
rather than holistic at that level.
In conjunction with the appeal to nonconceptual processes, Recogni-
tion theories typically posit a stage in the perception of space that is insu-
lated from the effects of conceptual knowledge and beliefs. Early visual
processes are cognitively impenetrable and domain specific, in the strict
sense described by Jerry Fodor (1983); that is, they are “modular.” More-
over, they depend on innate knowledge. This is a point that has sometimes
been applied specifically to pictorial space. In the words of Arthur Danto
(in press), who defends the modularity thesis: “As a pictorial device, per-
spective was invented by Brunelleschi, but as a mechanism of optical per-
ception [it] is genetically defined.” Techniques for representing space are
historical, he argues; but the spatial eye itself is not.
108 Mark Rollins

However, the relation of SDT to the modularity thesis is actually com-


plex. These complexities matter, I shall now argue, because it is precisely
with regard to the modularity thesis and the reliance on nonconcep-
tual foundations for perception that mainstream Recognition theories go
astray.

The Modularity Thesis


The primary problem with construing the directness of perception in
terms of the modularity thesis is that the thesis is probably false. At least
this is so with regard to the strong version advocated by Fodor. Some
suggestive evidence in this regard comes from neuroscience. For instance,
it has been shown that there are as many or more lateral and descend-
ing connections among functionally and anatomically distinct areas of a
monkey’s visual system as there are ascending connections (Van Essen
and Anderson 1990). Consistent with that fact is the evidence that the
responses of cells in area V4 to a visual stimulus can be modified by
somatosensory stimuli (Maunsell et al. 1991). Finally, recent brain-imaging
studies of humans show that the same areas of visual cortex that are acti-
vated by viewing a picture are also activated by remembering it. This sug-
gests that visual processes involved in picture perception, even basic ones,
can be affected by information stored in memory (Wheeler, Petersen, and
Buckner 2000).
The assumption is that such lateral and descending connections in the
brain serve some function in information processing. And the possibility
that they serve some function specifically in the perception of visual
representations is supported by experiments in psychophysics. There the
evidence shows, in a variety of ways, that “later” stage processes can affect
“earlier” ones. Consider, as an illustration, the fact that face recogni-
tion can influence (discernment of) shape-from-shading. A hollow mask
viewed from the back is often seen as convex rather than concave; that
is, its nose is perceived as pointing toward, rather than away from, the
viewer. This is so, even when the mask is illuminated from below rather
than above. Because such a change in illumination often transforms the
perception of convexity into an experience of concavity when the object
is not a face, it can be argued that the categorization of the object as a face
must be what blocks the expected effect of shading cues. That conclusion
appears to be borne out by a further test in which two identical masks are
placed side by side, with one inverted. The assumption is that it is hard to
see the inverted mask as a face; thus it should not override shading cues in
109 Perceptual Strategies and Pictorial Content

the way that an upright mask does. And this turns out to be the case (at
least up to a certain viewing distance): the inverted mask is seen as con-
cave rather than as convex. Thus, it has been argued that semantic cate-
gorization affects shape-from-shading (Churchland, Ramachandran, and
Sejnowski 1994, pp. 33–34). In which case, the strong modularity thesis
would appear to be false.
To be sure, Ramachandran has also argued that economizing strategies
presuppose the existence of modules, combinations of which get reinforced
for their success in performing a task and become part of an organism’s
standard repertoire. In that case, what the evidence actually shows is not
so much that the modularity thesis per se is false as that the nature of
modules and their impact on perception is not what Marr thought it to be.
Ramachandran suggests that there may be many such modules, over two
dozen in vision, whose functions are even more specialized than Marr’s
account would lead us to believe. But I am going to argue that even if the
modularity thesis is true, that fact does not matter much for our under-
standing of how pictures represent. That point applies with even greater
force, I believe, when there is the kind of proliferation of modules that
Ramachandran suggests. Such designer modules, as we might call them,
essentially undercut the original import of modularity, as providing a
fixed, basic set of distinct abilities that can contribute to the performance
of various perceptual tasks.
In any case, my claim is that modularity in any form is not able to
carry us very far toward a hybrid account and thus does not shed light on
the contribution of the direct component to pictures. The weakness of a
modularity-based hybrid account appears in two ways. First, the con-
straints imposed by modules are actually fairly minimal, to the extent that
tasks such as picture perception are performed on the basis of perceptual
experience, that is, conscious central processes on which diverse knowl-
edge is brought to bear. For example, although he defends the modular-
ity thesis, Danto’s view is predicated on the assumption that, especially
when perceived pictures are works of art, central processes must be brought
into play. In contrast to Gombrich, Danto believes that there is an inno-
cent eye. It exists in early vision. But for him, there is also a beguiled
and jaded eye, an eye of the world, which appears later in the process-
ing stream. And that is what eventually determines pictorial content.
Although picture perception depends on a basic perceptual competence
that all perceivers have in common, according to Danto, perceptual psy-
chology cannot tell us what the content, spatial or otherwise, of a picture
110 Mark Rollins

is. It cannot tell us what a picture means. The reason is that most of the
content of a picture is, in Danto’s words, “invisible.” It depends on his-
torical relations that cannot be seen, although they can be understood.
Danto’s point is that the content of a picture must be inferred; because
history, he says, supervenes on perception through the perceiver’s beliefs.
But if we agree that this sense is legitimate, then the modularity thesis
becomes more or less irrelevant. At higher levels, the constraints on mod-
ules can always be superseded. Danto suggests that the modularity of
mind insures that there will be something like theory-neutral observations
against which to check the visual hypotheses and interpretations that we
form when confronted with pictures. But if we think of the constraints
imposed by modularity on pictorial content as something like the con-
straints imposed by observation sentences on scientific theories, then it’s
clear that their role is not that of elements from which an interpretation is
constructed but only that of providing test cases for any interpretation.
And there may be many of those interpretations that can pass the test. If
so, then this stage-based compromise with Gibson is not very compelling.
In the end, it is the indirect component that carries all the explanatory
weight.
The second manifestation of the modularity-based hybrid’s weakness is
that constraints on vision really have more the status of antecedent causal
conditions than of premises, the conditions themselves having content, in
an inductive inference. And this points to the another reason that a stage-
based hybrid account falls short of being a genuine compromise with
direct realism. Even if the visual system were modular, the modules could
be harnessed together in a variety of ways in the performance of a given
perceptual task. That is, modularity does not preclude the possibility that
a wide variety of perceptual strategies might be employed in picture recog-
nition; in that sense, the relation between informational arrays and the
visual system cannot be one to one.

The Nonconceptual Nature of Spatial Perception


Some of the same problems found with the modularity thesis also arise in
regard to the nonconceptual nature of perceived space. Of course, if the
modularity thesis is false, that already casts doubt on the possibility that
perceptual content is wholly nonconceptual. If perceptual content is non-
conceptual, someone might argue, then it is not cognitively penetrable.
The reason is that “cognitive penetrability” is defined as the effect of
knowledge by way of inference, and knowledge is encoded in concepts.
Thus if the evidence shows that perception is cognitively penetrable, it
111 Perceptual Strategies and Pictorial Content

should follow that it is concept dependent.9 However, I think that it is


important to say more about the specific nature of the constraints that
nonconceptual spatial processes are supposed to impose on picture per-
ception. In general, the motivation for positing such processes is clear:
without them, either we must say (with Goodman) that pictorial and per-
ceived space are constructions that depend on acquired concepts and
learning, or we must say that they depend on innate concepts.
From the point of view of recognition theory, this amounts to a dilem-
ma. On the one hand, the learning hypothesis is unacceptable, because it
flies in the face of what Schier and others take to be the distinguishing
feature of picture recognition; what Schier calls “natural generativity.”
Essentially, the idea is that once a person has been able to recognize an
object or scene in one picture, that person can recognize—without any
further learning—objects and scenes in all pictures. For this to be the
case, presumably, that person needs some sort of foundational perceptual
process. Yet, on the other hand, the innateness hypothesis runs afoul of
the conceptual holism that Peacocke and others assume to obtain. That is,
if any perceptual categories are innate, they cannot be concepts, unless all
concepts are innate, because concepts are interdefined in a holistic net-
work. But if that is so, it will be impossible to individuate concepts (and
thus to ascribe meaning to words or pictures), because there would be no
fixed point of reference from which to begin.
The appeal to spatial aspects, scenarios, and the like is intended to be a
remedy for this dilemma. For example, Lopes develops an account of the
impact of spatial aspects on the rest of a picture’s content in terms of var-
ious forms of pictorial indeterminacy. Of special significance here is the
distinction between cases in which a picture simply leaves certain fea-
tures out and cases in which its spatial organization precludes them from
being represented. A picture will be noncommittal about the wrinkles and
bumps on the man’s face, if the artist simply omits them from the picture.
In that case, we understand that the man’s skin has texture, but nothing is
implied about their particulars. By contrast, if not all of the man’s nose is
visible because of the angle from which the artist has chosen to depict his
face, something is implied about its particulars, namely, the overall shape
and size of the nose. Thus we might say that the range of properties that
can be represented in a picture is determined—controlled is perhaps the
better word—by the spatial aspect it contains.
Moreover, these spatial aspects are supposed to play a central role in
the determination of the picture’s referent, where it has one, on the model
of demonstrative reference. The aspects themselves are defined by causal
112 Mark Rollins

relations between the picture producer and the objects represented in the
scene. These need not be actual causal relations; a painting might embody
a point of view that the artist would have, if that artist were to stand in a
certain place. But what the picture provides is a representation of the spa-
tial aspects of objects as defined by this perspective. In that case, the pic-
ture perceiver is able to gain access to the content of the picture by means
of this spatial aspect, which the perceiver can recognize without any
reliance on concepts.10 The net result is essentially a two-factored theory
of pictorial content: the reference of a picture is determined by psycho-
logical processes, the object of which is itself established by causal spatial
relations. The meaning or content of the picture in a larger sense is then a
function of higher-order interpretations that draw on background knowl-
edge, vested in a holistic conceptual scheme.
This is a theory of pictorial content that rests on shaky empirical
grounds. For example, as Patricia Churchland, Ramachandran, and
Terrence Sejnowski (1994) point out, although figure-ground organiza-
tion is generally thought to precede shape recognition, their model allows
for the order to be reversed and for shape recognition to contribute to
the identification of figure and ground. Other evidence also suggests
that the spatial organization of a picture does not always ground or con-
strain recognition in the way the Recognition theory seems to imply. For
instance, subjects appear to be able to spatially reorganize remembered
pictures and, as a consequence, identify new features and new objects in
them (Kosslyn 1994). Thus, it cannot be said that scenarios or spatial
aspects or perceptual reference frames are always fundamental. Even if
spatial properties are somehow encoded preconceptually, that fact does
not resolve the dilemma described here.
Moreover, what emerges from these results is that the processing of pic-
torial space and its relation to the recognition of depicted objects need not
be done in a single way; it depends on the strategies available to the per-
ceiver. For example, in some cases, subjects are not able to overcome the
initial perceptual organization of remembered pictures and so not able to
recognize new objects in them (Reisberg and Chambers 1991). However,
in those cases, the block can be eliminated if the subjects are allowed to
supplement perception or perceptual memory with support from motor
control. For instance, subjects can recognize verbal ambiguities from
memory (e.g., when the word “life” is repeated over and over in the mind,
it comes to be heard as “fly”); but only if allowed to engage in a kind of
silent speech. That involves more than simply rehearsing the sounds in the
113 Perceptual Strategies and Pictorial Content

“mind’s ear” without speaking. It also includes sending motor commands


to the mouth, tongue, and lips without moving those parts of the body in
noticeable or sound-producing ways. This is shown by the fact that sub-
jects cannot recognize verbal ambiguities from memory if they are chew-
ing gum. What this suggests is that the perceivers’ motor system works
together with visual or auditory memory in the performance of the task.
That is, perceivers exploit their strategic resources.

5.4 PERCEPTUAL STRATEGIES AND PICTORIAL SPACE

Let us return for a moment to the Recognition theorist’s idea that what a
picture represents depends on what can be seen in it, using ordinary per-
ceptual abilities, and to the attempt to reconcile that idea with the view
that pictorial art is somehow languagelike. Recall that the marriage of
perception and convention was supposed to be arranged by grounding
pictorial representation on pictorial reference, using demonstrative refer-
ence as a general model. The assumption was that spatial layouts relating
the scene to the position of a viewer are necessary to establish reference
and that these layouts are perceivable without benefit of concepts. Conse-
quently, their perception is not learned and not dependent on background
knowledge. But in light of the problems with the modularity thesis and the
nonconceptual process claim that I have cited, this marriage of perception
and convention is, I believe, on the rocks. So I now want to propose the
following change. Insofar as reference does constrain what a picture can
represent, the implication of SDT is that reference is established more by
strategic success in performing a task than by causal relations. What I
mean by “strategic success” is not just that a unique computational result
is obtained, or even that a goal is achieved or a target is hit. I mean, rather,
that establishing reference is done in an efficient way that conserves the
resources of the system. This implication is suggested by the various ways
in which SDT assigns important roles to attentional control.
For example, Churchland, Ramachandran, and Sejnowski (1994) argue
that attention shifts can precede eye movements in tracking a moving
object and thus bind the visual stimulus to a location. Although Church-
land and colleagues do not describe the mechanisms by which such covert
attention shifts might be carried out, an empirically well-grounded model
based on brain studies has been suggested by David Van Essen and
Charles Anderson (Van Essen, Anderson, and Olshausen 1994; Anderson
and Van Essen 1987). Further, Ramachandran and William Hirstein
114 Mark Rollins

(1999) argue that the manner in which the information in a picture is


processed is, to some extent, the result of innately defined attentional pref-
erences, which have been exploited by artists in various ways. In this
respect, picture perception is like the bottom-up location of diagnostic
features in nature; for example, the baby seagull’s recognition of its
mother’s food-laden beak by virtue of the stripes on the beak. Moreover,
this attentional “glue” is an essential ingredient in perceptual learning.
The acquisition of perceptual strategies, Ramachandran and Hirstein
claim, is driven to a large extent by the fact of limited attentional
resources. Successful perceptual grouping and binding are inherently
pleasurable (because of links between the visual and the limbic systems).
Thus we are rewarded for economy.
This implies that perceptual reference is grounded on perceptual strate-
gies that depend on attentional and other control mechanisms. In that
case, reference is not determined simply by where the gaze is directed (or,
with pictorial reference, the direction in which the picture is “looking,”
i.e., its point of view). Invoking attention in that way invites a theory
of reference based on causal relations (and nonconceptual, foundational
modes of vision). What SDT implies instead is that pictorial reference
depends on the type of attentional strategies it encourages its viewer to
use, together with the viewer’s own attentional “style,” that is, the dispo-
sition of her visual system to utilize its resources (modules, subsystems,
perceptual categories) in characteristic ways. For reference, SDT suggests,
is really referring: it’s something that people do through pictures and per-
ceptions. It is to that extent an achievement (as philosophers used to say),
so the explanation of it must include some account of how it gets done.
Goodman himself acknowledged the importance of eye movements in this
regard: “The fixed eye,” he said, “is almost as blind as the innocent one”
(1968, p. 287). But if Goodman’s idea is cashed out in terms of SDT, then
a picture becomes more like a speech act than a formal symbol: picture
perception, like perception generally, has an important performance
dimension. If art is like a language, it must be so in a different sense than
Goodman intended.11
What, then, does this discussion have to do with pictorial space? The
answer, as I shall now try to make clear, depends on defining the sense in
which SDT is partly direct. I have suggested that perception cannot be
direct by virtue of hardwired constraints and modularized operations; nor
is it direct just because it is, in its early stages, nonconceptual. Nonetheless,
I think that the emphasis on the role of spatial aspects and organization
115 Perceptual Strategies and Pictorial Content

in some Recognition theories provides a clue to how perceptual strategies


might play a similar role in constraining pictorial content. In the models
to which I have previously referred, shifts of attention are a matter of
moving the gaze across different regions of space; in that sense, spatial
operations contribute to the recognition of objects through the scanning
of contours or the location of diagnostic features. Moreover, insofar as
eye movements are involved, the process is itself identified with actual
changes in spatial position. This is true, by the way, even when the eyes do
not move. In the Van Essen and Anderson (1987) model of covert atten-
tion shifting, information is literally rerouted so that it flows along a dif-
ferent path in the brain.12 But beyond that, I think there is a sense in which
perceptual strategies contribute to the spatial content of pictures, and
their impact does not rest on inference of either the deductive or the induc-
tive type. To show how this is so, I turn now to an example of resource
deployment on a neural network model.
This example involves a theory of concept redeployment in connection-
ist terms. The account is motivated by the fact that connectionism does
not seem well equipped to explain visual discoveries—sudden changes in
the perceived identity of an object or, indeed, in the perceived spatial
organization of a scene. These changes cannot be easily attributed to the
gradual settling of a connectionist system into a stable state through
exposure to a training set over time. However, Paul Churchland (1989)
has tried to resolve this problem by proposing a model of concept rede-
ployment. Churchland’s original goal was to give an account of theory
change in science. But what he says can easily be extended to perceptual
learning and to the changes that define the history of visual art. The phe-
nomenon of the revelatory experience, the creation of a radically new way
of seeing things, is illustrated equally well by the physical theory that
treats light as a wave and the perceptual experience in which a figure first
seen as a rabbit is now viewed as a duck.
In both cases, what is distinctive is that no new concepts are added to
the scientist’s or perceiver’s store; old ones are redeployed instead. Thus
there is no learning in the most familiar sense: no new facts are discovered
(although unnoticed features may become apparent), and no new cate-
gories are created into which data are put. What makes this sort of con-
cept redeployment a candidate for direct perception is that it results from
a change in resource allocation and nothing else, just putting your percep-
tual resources in a different account. These revelatory experiences produce
no initial growth of intellectual capital, and thus no higher-order, complex
116 Mark Rollins

mental representations should be expected to ensue. Although it is easy to


think of the process as one of concept association, this is a case in which
changes in the strengths of connections among concepts should not be
construed as changing the basic conceptual scheme. Instead, the change
should be ascribed to control mechanisms, routing information along new
paths rather than representing it in new forms.13
I would argue, then, that Churchland has provided an alternative way
to understand the sense in which vision is partly direct, as exemplified by
this model of concept deployment. In so doing, he has pointed to a reso-
lution of a certain tension in his own account. According to this account,
the visual system encodes information in nonsymbolic, that is, nonlin-
guistic, forms. We might be inclined, then, to consider perception to be
direct simply by virtue of the fact. Indeed, some proponents of neural net-
work models, in which vision depends on connectionist principles, explic-
itly claim that this is the case. On these accounts, visual processes are
similarity based and subsymbolic. As a consequence, the models are said
to be Gibsonian in spirit. Although vision is mediated by internal repre-
sentations, the representations consist of activation patterns in a neural
network, for example, “perceptual prototypes” produced by vector coding
as parcellated regions in a multidimensional similarity space. In that
case, the principles of ecological resonance between an organism and the
environment are supposed to extend inward to the brain. In the words
of Bechtel and Abrahamsen (1991, p. 268): “The connectionist concept
of relaxing or settling into a stable state is reminiscent of the ecological
concept of resonating to an invariant.” Thus, “as cognitive psychology
becomes increasingly connectionist, it may become increasingly ecological
as well” (1991, p. 267).
The problem is that, under some construals at least, connectionist pro-
totypes are concepts, they are cognitively penetrable, and they are inter-
related in a holistic network. In all of these respects, they differ from the
representations of space on which recognition is supposed to depend.
Moreover, Churchland, for one, views connectionist vector coding as a
form of inductive inference. So to that extent, there can be no straight-
forward link between directness and nonsymbolic internal representa-
tions of a connectionist sort.
However, the tension in Churchland’s construal of perception as both
direct and inferential can be resolved by relocating the source of the
directness in his model. Directness in Churchland’s theory of perception
is to be found, not in the fact that vision depends on nonpropositional,
117 Perceptual Strategies and Pictorial Content

similarity-based representations but in the redeployments of perceptual


categories that I have just described. Those redeployments are driven by
attending to certain features of the visual stimulus, and they produce a
change in the “tuning” of the perceiver to his environment. Consequently,
they affect the “ecological laws” that might apply. In that sense, they con-
tribute to perceptual reference.
As it happens, this model of resource deployment embodies some of the
features that proponents of Recognition theories have found attractive.
For instance, what Churchland has in mind is a type of discovery that
precedes subsequent inductive reasoning. In that sense, a perceptual strat-
egy functions somewhat like a stage in information processing (although
not one that is universal or unmodifiable). Similarly, although concept
deployments obviously involve concepts, they also do not involve induc-
tive generalization, the formation of new concepts, or explicit thoughts in
propositional form. That is, they do not involve the type of concept use
that ordinary cognition requires. In that respect, we might speak of them
as being “nonconceptual.”
The account I have laid out also helps us understand some of the simi-
larities and differences among competitor accounts that are congenial to
the notion of perceptual directness. Specifically, the learning that SDT
envisions is often more a matter of conditioning or training (e.g., the train-
ing of attention) than of deriving conclusions from premises. This is appar-
ent in the emphasis of Churchland, Ramachandran, and Sejnowski (1994)
and Ramachandran and Hirstein (1999) on the role of reinforcement
learning in establishing perceptual strategies in the brain. This is a form of
learning they explicitly contrast with learning of a more cognitivist sort,
for which Marr’s constraints are said to provide a foundation. Thus it
may help distinguish their view from Cutting’s, for instance, which posits
a mathematical, deductive model for the direct specification of informa-
tion and allows for higher-order cognition in the perception of pictures of
certain types. In the latter case, which involves the elaboration of input
and the “building up” of constructions, learning in the sense of stored
background knowledge may be brought into play.14
Finally, this analysis of attention and other modes of resource control
sheds light on what is visible and what is not in a picture. Ramachandran
(1990) has suggested that there is a sense in which information that is not
explicitly encoded in perception is “captured” by motion and by attrac-
tion to diagnostic features. This suggests that although the information is
not represented, it remains somehow available, either for the performance
118 Mark Rollins

of a perceptual task or as a component of perceptual experience. A precise


correspondence between each spot on a leopard in different locations is
not computed, but information about the spots might be used to identify
the animal as a leopard, and perceivers report that they are aware of
changes in position of the stimuli that are captured in this way. Thus there
is a sense in which such information forms part of the content of percep-
tion but content that is unrepresented.
In this case, the strategy seems to involve ignoring some information for
the purpose of the task, while retaining it implicitly in the following sense:
the visual system assumes a rough correspondence between the informa-
tion that is represented and the information that is not. This assumption
allows the perceiver to experience the unrepresented information in some
sense, as well as perform tasks that depend upon it. And this is so, even
though the information is never explicitly encoded and thus cannot be
used to draw the inferences or make the computations needed to produce
more elaborate mental constructions. Although the notion of the capture
of information is not entirely clear, it invites comparison to the Recog-
nition theorist’s claim that spatial organization in a picture can ground a
kind of pictorial content that is implicit, that is, not simply left out but not
delineated either, because it is occluded or because it is not visible from the
point of view that the picture contains. I conclude, then, with some brief
remarks about how the bearing of perceptual strategies on the unrepre-
sented content of a picture.

5.5 THE UNREPRESENTED CONTENT OF PICTURES

Pictures are, of course, always incomplete. Details are inevitably omitted,


but also the point of view and spatial organization of the scene allow some
things but not others to be literally seen. It is tempting to say that this
is where indirectness intervenes: we fill in the gaps through inference or
imagination. This is the constructivist view. It is the sense in which picto-
rial art contains “invisible” content on Danto’s account. We add a layer
of conceptualization to our recognition of the type of item denoted by way
of a perspective and thus comprehend that there is more to the item than
literally meets eye. Indeed, according to this view, we comprehend what
more there is in a way that penetrates experience, while leaving the basic
elements of the perceived stimulus unchanged. This more inclusive experi-
ence then mediates the performance of the task. Our understanding that
certain features are likely to lie behind an occluding object in a picture is
119 Perceptual Strategies and Pictorial Content

the product of interpretation. It is constrained by what we see and, in turn,


limits our further thoughts about it.
However, I think that the unseen content of a picture can include more
than what lies behind the occluder. The reason is that we are constrained
in what we can see in front of the occluder as well. As Churchland’s exam-
ple of concept deployment suggests, patterns of information use tend to
become ingrained; they dominate and sometimes interfere with our dis-
covery of solutions to perceptual and scientific problems. Breaking their
grip often requires deviating from familiar routines or sequences in the
processing of information at different points in space; it requires attending
to stimuli in a novel way. The performance of a task becomes automatic,
and that may prevent us from noticing what is before our very eyes.
Moreover, insofar as pictorial content is the product of constraints, these
constraints are not construed as representations in the mind or brain. This
is as true of perceptual strategies as it is of assumptions that operate
in early vision. In that sense, too, perceptual strategies contribute to pic-
torial content that is unrepresented.
Finally, these patterns of control, these habits or predispositions, con-
tribute to the performance of a perceptual task in a larger sense than just
bringing the task to completion. These routines, relative to the perceiver’s
abilities and level of expertise as they are, affect the speed and efficiency,
as well as the reliability, with which the end is achieved. As a consequence,
they affect the perceiver’s experience: the sense of ease or effort, the feel-
ing of challenge or frustration, that accompanies the performance of every
task. Pictorial content, in this sense, too, is only implicit in picture per-
ception. Nonetheless, where the picture is a work of art, it permeates aes-
thetic experience. It is an important aspect of pictorial content that should
not be ignored.

5.6 CONCLUSION

It is sometimes said that the development of linear perspective reflects the


emphasis on the importance of the individual that was characteristic of
the Renaissance. It emphasized the centrality of the picture’s spectator,
who could comprehend all with little effort and in a single glance. It was
thus the natural complement to voyages of discovery and the conquest of
geographical space (Fleming 1986). That tells us little, of course, about
how linear perspective actually works and why (beyond the convergence
of everything from the picture’s point of view on a single vanishing point)
120 Mark Rollins

it highlighted the spectator. But it does suggest that pictorial space and the
appeal of perspective has something to do with mechanisms of perceptual
control. What I have argued is that adapting a plausible theory of picto-
rial representation, Recognition theory, to the various hybrid accounts of
perception that have emerged in recent years sheds light on how those
mechanisms operate. This light is, I think, more important than the ques-
tion of what it means for perception to be direct. But by articulating a
Recognition theory of this sort, we can also come to better understanding
of the differences among certain hybrid theories, which is where the most
interesting issues remain to be explored.

Notes
1. It will soon become clear that these are strategies available to the visual sys-
tem rather than the perceiver. I do not wish to imply by the term strategy that
optional modes of processing are always selected as a matter of conscious choice.
2. And in either this or its original version, the idea that there are varieties of pic-
tures can be combined with the pictorial surrogate view. The result would be that
pictorial surrogacy would then be said to come in a variety of forms.
3. Of course, we then have to define the term “indirect,” which is not a simple task
(as Schwartz 1994 has shown). I simply stipulate for now that “indirect” percep-
tion involves processes that depend on internal representations; that it does not
matter whether those representations are thought to be neural or irreducibly men-
tal (as perhaps a functionalist philosophy of mind suggests); and that to represent
something is to stand for it in some way. Thus perception is indirect if it depends
on an internal event that refers to the object perceived and if the attribution of
properties to that object presupposes the reference.
4. Compare also Stillings et al. (1987, p. 456).
5. However, on this account, pictures need not be perceived exactly as ordinary
objects are. There can be significant differences in the ways the standard percep-
tual equipment is employed.
6. There are significant differences among these thinkers. In particular, Peacocke’s
view is sometimes treated as opposed to a recognition theory of the sort that Lopes
endorses because Peacocke makes basic recognition depend on resemblance. But
these differences do not affect the point I wish to make here.
7. That is, an accurate understanding of pictorial reference by a perceiver is one
that comports with its actual causal history.
8. To this, Lopes adds an important qualification: A picture is not limited to a sin-
gle consistent point of view, as vision is. It may contain multiple and visually
impossible perspectives. Nonetheless, it is essential for Lopes to lay down some
constraints on the spatial organization of a picture and its aspects, if he is to be
able to distinguish the case of a single picture with multiple perspectives from the
case of multiple pictures side by side or in the same frame, each with a single point
121 Perceptual Strategies and Pictorial Content

of view. This he does in terms of “spatial unity.” A picture, he says, “is a repre-
sentation whose content presents a “spatially unified” aspect of its subject. . . .
[E]very part of the scene that a picture shows must be represented as standing in
certain spatial relations to every other part” (1996, p. 126). It is in this sense that
space plays a fundamental role for him. This would be true of cubist paintings or
impossible pictures, for instance, which contain multiple points of view but not
true of a picture postcard showing the various sights of Frankfurt. The former
could be related to a single object-centered representation; but the latter could not.
9. However, the converse is not true: perception might depend on concepts, but if
those are limited in scope (e.g., the concept of an edge, a line, an angle, etc. in early
vision) then perception need not be said to be affected by background knowledge.
Conceptualization does not entail cognitive penetration. This assumes, of course,
that concepts can be used in relative isolation, which is just the point that some
philosophers (Peacocke, for one) deny. To that extent, one’s view of the connec-
tions between concepts and cognitive penetrability in vision will depend on one’s
view of conceptual holism. In contrast to the other recognition theorists I have
cited, Schier does not seem to take a position on whether perceptual content, in
particular spatial content, depends on concepts or not. But for him, in either case,
early recognition processes are modular.
10. Of course, being nonconceptual is only a necessary and not a sufficient condi-
tion for spatial perception to play the dominant role that it is supposed to play. In
Marr’s model, shape and line representation involve no concepts either. Lopes
thinks that the perceiver-centered perspective is eventually to be translated into an
object-centered representation, as Marr has claimed. It is that translation, in fact,
that supports the recognition of objects under different aspects. Presumably, as
early stages of perception, shape and line recognition also constrain later stages.
Nonetheless, initially such information is viewpoint dependent; in that sense, spa-
tial aspects might be said to be fundamental.
11. Of course, the analogy to speech acts is only partial at best. The pragmatics
of language depend on more than the mechanics of verbal “gestures,” including
the intentions of the speaker and an audience as the primary context of speech.
But Recognition theory does not claim to explain all of reference in terms of
demonstratives, or even all of demonstrative reference in terms of spatial aspects.
My claim about perceptual strategies, attention, and referring is intended to be
equally modest.
12. It is important for the account I am developing here that such rerouting can
also occur in other ways that do not involve shifts of attention. For example,
brain-imaging evidence strongly suggests that after even a brief training period,
some perceptual tasks become virtually automatic, in the sense that they no longer
require attention. When that happens, they appear to be performed by different
areas of the brain than when the task was new (Posner and Raichle 1994, pp.
125–129).
13. Churchland’s model applies to resources other than concepts, for example,
representations of abstract shapes for which we have no labels, as might be pro-
duced by early visual modules. I omit here an account of the actual mechanism of
122 Mark Rollins

redeployment that Churchland suggests. It involves attaching different operators


to inputs or stored information to allow them to be used subsequently in different
calculations (Churchland 1992, p. 46). The attachment is not itself part of the cal-
culation, nor are the items to which the operators are attached changed when
different operators are used. This appeal to operators may seem to suggest that
“deployment” is a kind of inference after all, perhaps on the model of deduction:
a literal movement of shapes from one position to another, without regard for
their meanings or what they denote. However, the redeployments controlled by
the operators will not exemplify rules of inference, such as material implication or
negation. A better comparison might be to instructions for arranging pictures in a
gallery without regard for either what they represent or for relations among their
features, where the features are construed as object parts or governed by combi-
natorial principles.
14. It is important to note that what are reinforced on these SDT models are
strategies, that is, forms of overt behavior and neural response. What gets
rewarded is economy in resource use, not the truth of ideas or veridicality of per-
ception. The reward is for doing a task quickly, efficiently, and with little effort,
not ( just) for getting it right. Of course, accuracy in perception should bring fur-
ther rewards, and some level of accuracy will be a necessary condition for the
selection of a strategy in the species or the strategy’s acquisition by an individual.
But SDT suggests that we could be rewarded (at least internally) even when we get
things wrong, perhaps as a result of imposing a coherent organization on a stim-
ulus. Thus, the appeal to reinforcement learning by Churchland et al. does not
mean that theirs is a model of the “association of ideas” that might be treated sim-
ply as a form of inductive inference. Robert Schwartz (1996, p. 83) has argued that
Cutting’s account is hard to distinguish from theories that follow Ramachandran
(namely Gilden’s 1991). I agree that Cutting’s and Ramachandran’s approach
have important points in common. But I suggest that one important difference,
even if it is only a matter of emphasis, is the relative importance their views attach
to learning of different kinds.
PART II
THE STATUS OF PERSPECTIVE
This page intentionally left blank
Chapter 6
Optical Laws or Symbolic Rules? The Dual Nature of Pictorial Systems

John Willats

Two questions concerning the representational systems in pictures have


been widely debated. The first is whether perspective is a uniquely good
way of representing shape and space or whether it is just one among a
number of equally valid systems, rather as English is just one among a
number of languages. The second is whether representational systems are
natural or conventional. By “conventional” I do not just mean conven-
tional in the weak sense that it is or has been conventional to use perspec-
tive in some cultures but not in others,1 but in the stronger sense that the
relations between the spatial systems in pictures and the scenes they rep-
resent are arbitrary and can therefore be determined by cultural conven-
tions rather than natural laws.
In practice these two questions are often confused, because it is often
implied that perspective is the best of all possible systems precisely because
it is natural—that is, based on the laws of optics. The obvious model here
is photography. Because cameras can produce pictures without the inter-
vention of human agency,2 and because these pictures are necessarily in
perspective, it is tempting to think that perspective must be the best of all
possible systems. The problem with this is that it suggests that other pic-
torial systems, especially relatively unfamiliar ones used in other periods
and cultures, must somehow be inferior. Much of this confusion can be
avoided by asking what it is that perspective is good for, and I have argued
elsewhere (Willats 1997) that different representational systems are good
for different purposes. Perspective, for example, can be good for land-
scape painting because it gives one a good idea of what a landscape looks
like from a particular viewpoint, but orthographic projection is better for
engineering drawings—and is universally used for that purpose—because
it shows the true shapes and sizes of objects, irrespective of any partic-
ular viewpoint. Similarly, railway maps based on topological geometry,
126 John Willats

such as the map of the London Underground system, are better as route
maps than pictures in perspective because the maps show the connections
between stations without irrelevant details such as the shapes of the lines
or the stations.
There is, however, another more subtle source of confusion concealed
in these questions: the temptation to judge representational systems as
natural or conventional on the basis of whether they are best described in
terms of projective geometry or symbolic rules. “Natural” theories of pic-
tures are derived from the laws of optics, and because the projection of
light rays from the scene and their intersection with the picture plane can
be described in terms of three-dimensional projective geometry, the argu-
ment that pictures are natural has focused on the question of whether the
spatial layout of pictures can be described in these terms. Margaret Hagen
(1985, 1986) has argued that all representational pictures can be regarded
as being based on what she calls “natural perspective” because their spa-
tial systems can be described in terms of three-dimensional projective
geometry. In contrast, other types of pictures such as Byzantine mosaics
and cubist paintings, which are based on inverted perspective and whose
spatial systems are difficult to describe in these terms, are often said to be
“languagelike” because they symbolize features of the scene rather than
represent them optically (Goodman 1968). This analogy is typically used
to support the argument that pictures are conventional because linguists
generally agree that the relations between the units of language and their
referents are arbitrary. As a result, it is tempting to conclude that in the
case of pictures whose spatial systems are best described in terms of sym-
bolic rules, the relation between scenes and pictures is also arbitrary. My
thesis here is that these two modes of description—in terms of the laws of
optics or symbolic rules—are not mutually exclusive but complementary
and that in many cases descriptions given in terms of symbolic rules can
also be related back to optical laws. However, I shall also argue that in
some circumstances one form of description may be more appropriate and
revealing than the other.
Take, for example, Giovanni Canaletto’s line drawing in perspective of
the Campo di Santi Giovanni e Paolo, circa 1735.3 Martin Kemp argues
that the spatial system in this and a number of similar drawings “is best
explained by supposing that it is based on a camera image” (1990, p.
197)—that is, that Canaletto produced these drawings with the aid of a
camera obscura, the forerunner of the modern photographic camera. To
support his argument Kemp draws attention to the unusual quality of the
127 Optical Laws or Symbolic Rules?

line in this drawing: “[I]t traces the perspectival outlines in a summary,


freehand manner with virtually no auxiliary construction lines” (1990, p.
197). What Kemp is saying is that Canaletto looked directly at the optical
image in his camera obscura and then traced over the outlines of the build-
ings as they appeared in the image. This method of producing a drawing
was not unusual in the eighteenth century. However, pictures in perspec-
tive can be and were produced by quite other means: that is, by the use of
perspectival rules, such as the rule that lines representing edges in the
scene at right angles to the picture plane (the orthogonals) converge to a
vanishing point. If Canaletto had used this rule in the production of the
Campo di Santi Giovanni e Paolo, Kemp suggests, the construction lines
derived from this rule would probably still be present in the drawing.
Such rules have a long history. The first known example of a construc-
tion rule resulting in a picture in true perspective was given by Alberti in
his Della Pittura written in 1436 (Spencer 1966) in the form of what he
called the costruzione legittima, but this replaced earlier workshop rules,
such as the rule that the depth of each successive tile in the representation
of a pavement should be reduced by one-third, or Cennino Cennini’s rule
that the moldings at the top of a building should slope downward while
the moldings at the bottom should slope upward. Following these earlier
rules results in pictures in which the orthogonals converge but not to a sin-
gle coherent vanishing point. In practice, artists often combined optical
methods with the use of perspectival rules: it seems likely that Canaletto
would have corrected the composition in his painting of the Campo di
Santi Giovanni e Paolo by using perspectival rules,4 even though the
original outlines may have been taken from a camera obscura image. In
the case of true Renaissance scientific perspective the rules described by
Alberti are directly related to the laws of optics. Nevertheless, it may be
appropriate to describe the actual processes of picture production for pic-
tures such as the Campo di Santi Giovanni e Paolo, in terms either of the
laws of optics or the use of symbolic rules, depending on which method,
or combination of methods, the artist actually used.
When it comes to picture perception, descriptions based on optics
and three-dimensional projective geometry are usually more appropri-
ate, because it is characteristic of pictures in perspective that their two-
dimensional geometry is similar to that obtained by projecting the rays of
light from a real scene on to a flat surface. Again, definitions of this kind
have a long history, from Brook Taylor’s (1719) definition of a “Picture
drawn in the utmost Degree of Perfection” to James J. Gibson’s definition
128 John Willats

of a “faithful” picture as “[a] delimited surface so processed that it yields


a sheaf of light-rays to a given point which is the same as would be the
sheaf of rays from the original scene to a given point” (Gibson 1954;
quoted in Gibson 1971, p. 28). In other words, pictures in perspective
provide us with a view that, at least as far as its geometry is concerned, is
similar to what we might obtain from a view of a real scene. On the other
hand, we characteristically recognize pictures as being in perspective (as
distinct from being in some other system) in one of two ways: either by
looking to see whether their orthogonals converge to a vanishing point
(in the case of pictures depicting buildings or other rectangular objects) or
by looking to see whether objects that are farther away in the scene are
drawn to a smaller scale compared with objects in the foreground (in the
case of landscapes in which there are no obvious straight edges). Here
again, as in the case of picture production, descriptions in terms of the
laws of optics or symbolic rules may be more or less appropriate, depend-
ing on the circumstances.
Similar arguments apply to pictures in systems other than perspective.
Perspective is just one among a number of drawing systems that can be
defined in terms of either drawing rules or three-dimensional projective
geometry: these include oblique projection, the basis of many oriental
drawings and paintings, and orthogonal projection, widely used by engi-
neers and architects for their working drawings. In practice, engineers
and architects do not derive their working drawings from optical views of
objects or scenes but make use of implicit rules of the kind “draw the front
faces as true shapes” (for drawings in orthogonal projection) or “draw
the side edges using parallel oblique lines” (for drawings in oblique pro-
jection). Similarly, although we know little of the working methods of
Chinese, Japanese, and Persian artists, all of whom used oblique projec-
tion, it seems most unlikely that they used optical methods: in the vast
majority of cases real scenes corresponding to these pictures did not exist,
and there would have been insuperable difficulties obtaining a view from
an infinite, or at least very considerable, distance above the scene, which
is what definitions in terms of optics demand. It seems much more likely
that these artists made use of implicit symbolic rules (Willats 1997, pp.
198, 199). However, when it comes to picture perception, definitions of
these systems based on projective geometry seem more appropriate because
all these systems provide at least approximations to views of possible
objects or scenes.
Hagen has argued that because oblique projection, orthogonal projec-
tion, and their variants can be described in terms of three-dimensional
129 Optical Laws or Symbolic Rules?

projective geometry, pictures based on these systems exploit what she calls
“natural perspective, the geometry of the light that strikes the eye” (1986,
p. 8). To the extent that such pictures provide at least approximations to
views of possible views of objects or scenes this argument has some force
but only in relation to picture perception. However, her failure to recog-
nize that these systems can equally well be defined in terms of symbolic
rules—rules about the lengths and directions of the orthogonals—has led
her to a number of false conclusions.
The first of these is her claim that “there is no development in art”
(1985, p. 59), which she applies to both artists’ pictures and children’s
drawings. She was obliged to make this claim because it follows from her
contention that “all representational pictures from any culture or period
in history” and all children’s drawings (beyond the scribbling stage) are
based on “natural perspective, the geometry of the light that strikes the
eye” (1986, p. 8). If oblique projection, orthogonal projection, and the like
are simply varieties of perspective, and perspective is defined in terms of
three-dimensional projective geometry, how can there be any develop-
ment from one system to another? As Hagen puts it, “is it possible to
measure developmental level with a tool that itself shows no develop-
ment?” (1985, p. 76). The flaw in Hagen’s argument is that some periods
in art history do show a definite developmental sequence, and this is also
true in the development of children’s drawings. Moreover, this sequence is
not difficult to understand once the projection systems are defined in terms
of drawing rules.
Perhaps the most obvious example is the case of children’s drawings
of rectangular objects such as tables or cubes. Although the details vary
somewhat according to the experimental conditions, and the terminology
varies from one writer to another, a number of experiments have shown
that children’s drawing development follows the sequence: topological
geometry, orthogonal projection, horizontal and vertical oblique projec-
tion, oblique projection, and, finally, some form of perspective (Lee and
Bremner 1987; Nicholls and Kennedy 1992; Victoria 1982; Willats 1977).
This sequence is easily described in terms of increasingly complex draw-
ing rules. Thus in drawings based on topological geometry there are no
orthogonals; in orthogonal projection, the orthogonals are represented by
points; in horizontal and vertical oblique projection, the orthogonals are
horizontal or vertical; in oblique projection, the orthogonals run at an
oblique angle across the picture surface; and in perspective, they are also
oblique but converge to a vanishing point. So the rules for producing
pictures in these systems are increasingly complex (Willats 1977; 1997). In
130 John Willats

this case, describing children’s drawings in terms of drawing rules pro-


vides a convincing interpretation of the developmental sequence, whereas
describing them only in terms of projective geometry leaves this sequence
inexplicable.
The reasons this sequence occurs are perhaps more controversial, but
one factor seems to be that the acquisition of more complex rules is nec-
essary in order to produce drawings that are better as representations, in
the sense that the objects depicted can more easily be recognized and their
three-dimensional shapes can more easily be seen in the drawings (Willats
1997; Wollheim 1977). This argument, however, is best related to picture
perception and definitions of the systems given in terms of projective
geometry: it is only by the employment of systems later in the develop-
mental sequence, such as oblique projection and perspective, that draw-
ings can be produced that are better as representations, because they show
rectangular objects from a general direction of view, and false attach-
ments can be avoided. A “false attachment” in a picture occurs when
two features in a scene appear to be touching, whereas in fact they are
separated in depth: for example, when in an amateur photograph a tree
appears to be growing out of someone’s head.
A similar sequence occurred in the development of Greek vase paint-
ing. Again, the terminology varies from one writer to another. In their
drawings of rectangular objects the development followed the sequence
of orthogonal projection, oblique projection, and, finally, a simple form
of perspective that developed into more or less true perspective for wall
paintings during the Hellenistic period (White 1967; Willats 1997). As
with the development of children’s drawings, this development is prob-
ably best understood in terms of the acquisition of increasingly complex
drawing rules, at least so far as picture production is concerned; the
impulse behind this sequence was probably, again, the desire to produce
drawings that were better as representations. Finally, the change from the
use of inverted perspective or incoherent mixtures of oblique projection or
naive perspective during the medieval period to the discovery or rediscov-
ery of true perspective at the beginning of the Renaissance was regarded
at the time as a definite advance; and as Alberti made clear, this advance
was achieved by the use of better rules.
Hagen’s failure to recognize that the spatial systems in pictures can also
be described in terms of drawing rules leads her to a second false conclu-
sion: the claim that all representational pictures exploit natural perspec-
tive. Again, this is mistaken. There are two quite large classes of pictures
131 Optical Laws or Symbolic Rules?

whose representational systems cannot be defined in terms of the laws of


optics but which are, nevertheless, representational in any normal sense of
the word. The first class consists of pictures based on topological rather
than projective geometry: this class includes many cartoons and cari-
catures, children’s early drawings such as the so-called tadpole figures,
early rock drawings, the earliest Greek vase paintings of humans and
animals, and some artists’ pictures—among them paintings by Henri
Matisse, Picasso, and Paul Klee (Willats 1997). A good example is Klee’s
Another Camel! 1939.5 It is difficult, though perhaps not impossible, to
regard drawings of this kind as projections from views of scenes (Bunge
1966, p. 221); they can, however, easily be described in terms of topologi-
cal geometry. The shapes of the lines in this drawing are wildly distorted,
judged in terms of the laws of optics, but Klee has been very careful to pre-
serve the topological connections between the line junctions that denote
points of occlusion. As a result, the drawing is easily recognizable as a
depiction of a camel, and the three-dimensional shapes of the features in
the drawing can still be seen, albeit in a distorted form.
The second class of pictures whose spatial systems are difficult to
account for in terms of three-dimensional projective geometry are those
in inverted or divergent perspective and pictures containing incoher-
ent mixtures of systems, usually variants of oblique projection. This
class includes medieval paintings, Persian miniature paintings, Byzantine
mosaics, Russian icon paintings, and many modern paintings, includ-
ing cubist paintings. Hagen claims that inverted perspective: “was never
adopted anywhere with enough consistency to be called a style of spatial
depiction. . . . Whatever the reasons for the occasional appearance of
divergent perspective, it is never characteristic of any coherent art style”
(1986, p. 149). This is not, however, the case: although inverted perspec-
tive is not in itself a consistent system in the way that scientific perspective
is a consistent system, its employment in Orthodox art was consistent over
a period of at least a thousand years.
As with oblique projection and its variants, the spatial systems in all
these types of pictures can easily be described in terms of the rules gov-
erning the directions of the orthogonals. In inverted perspective, for exam-
ple, there is a general tendency for the orthogonals to diverge, instead of
converge, as they do in scientific perspective. However, these spatial sys-
tems cannot be regarded simply as alternatives to scientific perspective, in
the way that English and French can be regarded as alternative languages.
The human visual system is designed to take in spatial information from
132 John Willats

scenes, not from pictures, and judged by the standards of optical realism
pictures in these systems must be regarded as anomalous. But what could
be the reason for the existence of such anomalies?
There appear to be at least four possible reasons for these anomalies.
The first, and simplest, is the use of crude or mistaken rules or of incom-
patible mixtures of rules. For example, the fifteenth-century Birth of
Saint John the Baptist by Giovanni de Paolo is in a crude form of per-
spective but also contains incompatible mixtures of oblique projections.6
Anomalies of this kind often occur during periods of transition when
one drawing rule is being replaced by another: in this case, when oblique
projection and its variants were being replaced by perspective. Similar
anomalies appear in children’s drawings, during the transitions between
different developmental stages.
Second, as I have argued elsewhere, the anomalies that are so charac-
teristic of Byzantine mosaics and fifteenth- and sixteenth-century Russian
icon paintings are not “mistakes” but may have served quite specific the-
ological purposes. Unlike Catholic fifteenth- and sixteenth-century art,
and seventeenth-century Protestant art, Orthodox paintings and mosaics
were not intended to look optically realistic; rather, they were intended to
depict a spiritual world, as opposed to the material world of the senses.
Moreover, the flattening of the pictorial space that resulted from the use
of such anomalous devices as inverted perspective and false attachment
drew attention to the picture as a physical object so that the union of the
depicted spiritual world and its material support provided a metaphor for
the incarnation (Willats 1997).
Third, many artists, particularly during the twentieth century, used
anomalies for expressive reasons. Much of the poetic quality of Marc
Chagall’s paintings, for example, can be attributed to the fact that they
contain mixtures of drawing systems that impart a dreamlike and other-
worldly quality to the painting, similar to that found in many icons, and
it is surely no coincidence that Chagall was influenced by icon painting.
Finally, a number of modern painters have used anomalies as a way of
investigating the nature of painting itself. This seems to have been the
impulse behind much of the work of such diverse painters as Juan Gris,
perhaps the purist and most technical of all the cubist painters; the surre-
alist painter René Magritte; and Paul Klee. An adequate account of the
anomalies in the work of these painters, however, requires a fuller account
of the spatial systems in pictures than that given earlier in terms of topo-
logical and projective geometry.
133 Optical Laws or Symbolic Rules?

According to David Marr (1982) any scheme for representing shape and
space (and this includes pictures as well as scenes) must have at least two
components. The first component defines the spatial relations between the
units of which the scheme is composed. For example, the shapes of the
actual buildings in the Campo di Santi Giovanni e Paolo in Venice might
be described in terms of the relative directions of their edges in three-
dimensional space, whereas in drawings of this scene their shapes would
be represented by the relative directions of the lines representing those
edges. I shall refer to the systems that map the spatial relations in scenes
into corresponding relations on the picture surface as the drawing systems.
The second component in Marr’s account consists of the units of which
such representations are composed: Marr calls these the primitives of the
system. For example, a description of the shapes of the buildings in a
scene might be given in terms of edges as scene primitives, and the repre-
sentations of these buildings in a drawing might be made up of lines as
picture primitives. I shall refer to the systems in pictures that map scene
primitives into corresponding picture primitives as the denotation systems.
Thus an account of the representational systems in pictures can be given
in terms of two mutually dependent components: the drawing systems and
the denotation systems. In the drawing systems
spatial relations are mapped by the spatial relations
in the scene drawing systems into in the picture
and in the denotation systems
scene primitives are mapped by the picture primitives
(such as edges) denotation systems into (such as lines)
There are five main drawing systems: perspective; oblique projection
and its variants; orthogonal projection; inverted perspective; and systems
based on topological geometry. In addition, there are three main denota-
tion systems: optical systems; line drawings; and silhouettes in which the
primitives are two-dimensional regions. As (with certain restrictions) any
of the drawing systems can be combined with any of the denotation sys-
tems, this provides quite a rich method of classification that can be used
to describe the representational systems in pictures from a wide variety of
periods and cultures.
As with the drawing systems, the denotation systems can be defined in
terms of either the laws of optics or symbolic rules. In optical systems
zero-dimensional dots of tone or color are used to denote the intercepts of
small bundles of light rays. Typical examples are television pictures and
134 John Willats

pointillist paintings. The denotation system in Canaletto’s painting of the


Campo di Santi Giovanni e Paolo can be assigned to this class, although in
this case (as in most color photographs) the dots of color are extremely
small. In line drawings the picture primitives are one-dimensional lines—
normally black lines on a light ground—and examples can be found in
virtually all periods and cultures. In the third main class of denotation sys-
tems the picture primitives are two-dimensional regions. Typically, in pic-
tures based on regions as primitives, the regions denote whole volumes in
the scene, although they can also be used to denote regions in the visual
field. Regions and volumes have very few shape properties, the most
important being that of extendedness: that is, the property of relative
extension in various directions in space. In children’s early drawings, for
example, such as the tadpole figures, round regions are used to denote
round volumes such as the head or head/body, and lines or long regions
are used to denote long volumes such as the arms and legs.7 Similarly, in
the map of the London Underground, round regions are used to denote
the stations and long regions are used to denote the track. Examples of
other pictures employing regions as primitives include early rock draw-
ings, such as those of Camonica valley, and some paintings and drawings
by Matisse and Picasso (Willats 1997).
Because the tones and colors of the dots of pigment in a photograph or
television picture can be directly related to the fall of light, it seems appro-
priate to describe the denotation systems in these pictures in terms of
optics. But can this approach also be applied to line drawings? The prob-
lem here is that the array of light given off from a line drawing is usually
quite different from that which could be obtained from a real scene.
Compare, for example, Canaletto’s original sketch of the Campo di Santi
Giovanni e Paolo and his painting of the same subject. The array of light
given off from a painting, even a highly finished painting such as this, can
never fully match that given off by the scene it represents, if only because
the range of luminance levels in a painting can never be as great as that
obtained from a real scene. Nevertheless, the tones and colors in this
painting are so clearly related to the fall of light in the real scene that they
are better described in terms of optics than symbolic rules. But, equally
clearly, this is not true of the lines in the original sketch: most of these
lines represent the edges of buildings, edges in what Marr (1982) would
call an object-centered description of the scene that exist independently
of the fall of light. Thus, although Canaletto may well have derived the
drawing system in this picture from a camera obscura image, the deno-
135 Optical Laws or Symbolic Rules?

tation system would have been based on an implicit rule: “use lines to
denote edges and contours.”
Gibson maintained that “there is no point-to-point correspondence of
brightness or color between the optic array from a line drawing and the
optic array from a scene” (1971, p. 28). Certainly, early attempts to obtain
line drawings from gray-scale images automatically, by scanning photo-
graphs and representing the luminance steps in the image (points where
the tonal values change abruptly), are fairly unconvincing (Marr 1982).
However, research by Don Pearson, Hanna, and Martinez (1990) suggests
that Gibson’s assertion that there is no correspondence between arrays
from line drawings and real scenes may have been unduly pessimistic,
because they were able to produce quite respectable line drawings from
gray-scale images by using what they called a “cartoon operator” tuned
to pick out a combination of luminance steps and luminance valleys.
Nevertheless, what works for a computer does not necessarily work for
human beings. Many of the mistakes made by amateur artists and art stu-
dents come about as a result of the temptation to use lines to represent the
abrupt changes in luminance values that come about at the edges of shad-
ows, and this is particularly the case in figure drawing. An important part
of an artist’s training consists of learning to ignore the boundaries of the
shadows that often mask the points where contours end and, instead, to
use the implicit rule “use lines to denote contours.”
Thus when it comes to picture production, the denotation systems in
optically realistic pictures are often best described in terms of optics and
the fall of light, whereas the denotation systems in line drawings are usu-
ally best described in terms of symbolic rules. But if so, why is it that line
drawings can look so convincingly realistic? Pearson and colleagues sug-
gest that this might be because the early stages of the human visual sys-
tem include a feature detector, similar to their cartoon operator, that picks
out a combination of luminance step edges and luminance valleys, so
although the optic array from a line drawing is not physically similar to
that obtained from a real scene, it nevertheless provides an effective per-
ceptual equivalent. If this is indeed the case, it might make sense to
describe the perception of line drawings in terms of these features of the
optic array, rather than in terms of rules about what lines may and may
not denote.
Finally, are the denotation systems in pictures based on regions as
primitives best described in terms of optics or symbolic rules? It is not
difficult to see how pictures of this kind can be derived using symbolic
136 John Willats

rules, such as the rule “use a round region to denote a round volume,” and
it seems likely that children employ implicit rules of this kind as a basis for
their earliest drawings. However, it also seems likely that rules of this kind
have a “natural” basis in optics, because round volumes will always proj-
ect a round region in the visual field, whereas a long volume will usually
project a long region, except in the unlikely case in which a long region
is viewed end-on. In evolutionary terms, such “natural” associations pre-
sumably come about as the result of an association between the shape of
the object and the view of that object that is most frequently encountered
(Willats 1992b). Thus William Stern, commenting on the lines and round
regions that children use in their tadpole figures, said: “We call these sym-
bols ‘natural’ because their meaning does not first require to be learnt (as
in the case of letters or mathematical signs) but directly occur to the child,
and are used by him as a matter of course. Thus a long stroke is a natural
symbol for an arm or leg, a small circle for an eye or head” (1930, pp.
369–370).
By the standards of optical realism and projective geometry, tadpole
drawings must be judged to be anomalous, and as adults we do in fact
make such judgments when we look at children’s early drawings—which
is perhaps why they look so puzzling. When we look at tadpole figures we
interpret the lines as representing contours, and these lines look “wrong”
as the projections of the contours of the human figure. For the child, how-
ever, the lines do not represent contours, but form the outlines of regions
representing whole volumes. Thus the drawing rules underlying the per-
ception of these drawings by adults are quite different from the rules
by which such drawings have been produced. Perhaps an analogy might
be drawn here between children’s early drawings and their early speech.
Although sentences such as “Nobody don’t like me” sound wrong to
adults, “It is important to understand that when children make such
errors, they are not producing flawed or incomplete replicas of adult sen-
tences; they are producing sentences that are correct and grammatical
with respect to their current internalized grammar” (Moskowitz 1978,
p. 89). I suggest that, in the same way, the drawing rules by which the
tadpole figures have been produced are consistent in their own terms;
but they are not the rules normally used by adults.
According to Breyne Moskowitz, “Children’s errors are essential data
for students of child language because it is the consistent departures from
the adult model that indicate the nature of a child’s current hypotheses
about the grammar of a language” (ibid., p. 89). Similarly, the anomalies
137 Optical Laws or Symbolic Rules?

in children’s drawings can give important information about the nature


of the drawing rules they are using at each stage in the developmental
sequence. The use of anomalous or ungrammatical sentences as a way of
investigating the normal adult rules of language is, of course, a standard
technique in linguistics, made famous by Noam Chomsky (1957/1971),
and a similar technique was used in the so-called picture grammars devel-
oped in the 1970s by Max Clowes (1971) and David Huffman (1971), and
later by David Waltz (1975). In these grammars the “alphabet” is made up
of various types of line junctions, and the “syntax” consists of drawing
rules about the ways in which these junctions may and may not be re-
lated. These rules are not, however, arbitrary, but are derived from con-
straints about the nature of the physical world or the laws of optics. One
important rule, for example, is that a given line must have the same mean-
ing along its whole length so that a line cannot represent a convex edge at
one end and a concave edge at the other. Another rule about the num-
ber of surfaces hiding a contour is derived from the laws of occlusion, in
which one object may hide another from a particular direction of view.
Clowes and Huffman showed that line drawings that satisfy these con-
straints look “right,” while drawings that do not satisfy them look anom-
alous or “wrong.”
Well before the 1970s, however, a number of artists seem to have been
engaged in a similar enterprise, using anomalies in painting to investigate
the nature of painting itself. Magritte’s The King’s Museum, 1966,8 for
example, contains a very obvious anomaly, and in this case it is fairly
easy to see which pictorial rule is being broken. There is nothing odd
about the drawing system in this painting, which is straightforward per-
spective. What is odd is that in one part of the picture the normal rule of
occlusion is reversed. This rule says that parts of a scene will be hidden
from a viewer at a given point, and that representations of these parts of
the scene should be omitted from the picture. In the vast majority of pic-
tures, whether in perspective or otherwise, obeying this rule is so fun-
damental that we simply take it for granted: as Jan Koenderink (1990)
remarked, the representation of occlusion is the most powerful of all
depth cues. What Magritte’s painting does is to make this rule explicit.
Because the normal rule has been broken it has to be formulated in terms
of a denotation rule, but in this case the rule is clearly related to the laws
of optics.
The main drawing system in Gris’s Breakfast, 1914 is axonometric pro-
jection (a variety of oblique projection popular with the cubists), but this
138 John Willats

is disguised with diverging lines that give an impression of inverted per-


spective, and it is combined with a reversed version of atmospheric per-
spective in which the warmest and most saturated area of color occurs in
the background of the depicted scene.9 These drawing systems are com-
bined with no fewer than four anomalous denotation systems. First, as in
Magritte’s The King’s Museum, the normal rules of occlusion are reversed:
parts of the scene that would be hidden from the viewer are included, and
other parts that would be in view are left out. Second, there are numerous
instances of false attachments both within the picture and between fea-
tures in the picture and edges of the frame, breaking Huffman’s constraint
that the scene should be depicted from a general position of view. Third,
real objects in the scene, such as wood grain paper and real wallpaper,
are included in the picture, confounding the distinction between scene
primitives and picture primitives. And finally, by using part of a real news-
paper in the painting, Gris introduced an alien symbol system—natural
language—into the domain of the picture. These last two devices are
purely artificial, but otherwise all the drawing systems and denotation sys-
tems in this picture can ultimately be related back to the laws of optics,
although apart from the use of axonometric projection they reverse the
normal rules derived from these laws.
Much of Klee’s work, particularly in his later years, sprang from a con-
cern with the rules of depiction. As he said in his diaries: “ I have carried
out many experiments with laws and taken them as a foundation. But an
artistic step is taken only when a complication arises” (Klee 1961, p. 454).
Some of these “complications” seem to have been used for expressive pur-
poses, as in his Naked on the Bed, 1939.10 In other paintings and drawings
they seem to serve a purely investigatory or metalinguistic function: apart
from its playful title, his Oh! but oh! 193711 could be an essay on the deno-
tation systems in line drawing in the manner of Clowes or Huffman, and
in other paintings and drawings, such as With Green Stockings, 1939, the
two functions seem to have been combined.
With Green Stockings contains four types of pictorial anomalies, all
related to the denotation systems in the picture. The first group of anom-
alies concerns the way the girl’s arms, legs, and nose are drawn. In normal
line drawings the lines denote edges and contours, as they do here in the
outlines of the girl’s head, her skirt, and the ball. In other parts of the pic-
ture, however, the lines denote either three-dimensional volumes, as in
the case of her arms and legs, or a ridge (a two-dimensional surface) in
the case of her nose. In the painting (figure 6.1A and plate 3) the anoma-
139 Optical Laws or Symbolic Rules?

lous denotation systems used for the arms and legs are highlighted with
patches of color: yellow for the arms and green for the stockings. In fig-
ure 6.1B these areas are shown using two different tones. Systems of this
kind are very rare in adult pictures but common in children’s early draw-
ings such as the tadpole figures.
The second group of anomalies violates Huffman’s rule that a line must
keep the same meaning along its whole length: in Klee’s painting this rule
is broken in at least six places. One instance of this can be seen in the line
that begins in the top of the patch of red paint representing the girl’s hair.
In its path round the girl’s head this line begins by representing the con-
tours of her head, cheek, and chin; then it turns sharply and continues as
a representation of her shoulder. Just past this sharp turn, however, at a
point marked by the beginning of the patch of yellow paint (1), the line
changes its reference and denotes the girl’s arm as a long volume. Two
more anomalous changes of reference occur in the line that begins at the
top right of the painting. At first this line represents the girl’s arm as a long
volume. Then, just past the edge of the patch of yellow paint it turns and
changes twice, representing first the back of her blouse as a contour (2)
and then the line of her belt as a long region (3). Finally, no fewer than
three anomalies are present in the near-vertical line in the center of the
painting. At the bottom of the painting, within the patch of green paint,
this line represents her foot and her leg as long volumes. Then, as it
crosses the edge of her skirt (4), it represents the contour of a fold in her
skirt seen against the skirt as a background. Next, as it passes the line
of her belt (5) it changes its reference again, representing the contour of
her bodice, but this time against a blank background to the left. Finally,
the line changes its reference for a third time (6), and ends by representing
the ridge in the girl’s neck as she turns her head.
The third group of anomalies in the picture concern the way Klee uses
line junctions. In normal line drawings of smooth objects, there are two
types of line junctions: T-junctions, which represent the points where con-
tours disappear behind surfaces, and end-junctions, which represent the
points where contours end. With Green Stockings contains examples of
both types of junction, used both anomalously and in the normal way. A
normal T-junction occurs at the bottom right of the painting, where the
line representing the girl’s left leg disappears behind the edge of her skirt.
A similar T-junction occurs at the bottom left of the painting, but here the
line of the leg, instead of disappearing behind the edge of the skirt, con-
tinues in view. We may interpret this anomalous junction in two ways.
140 John Willats

Figure 6.1
(A) Paul Klee, Mit Grünen Strümpfen (With Green Stockings), 1939, watercolor
and blotting paper, 34.9 cm _21.0 cm (13 43 ˝_ 841˝), Felix Klee Collection, Berne.
© 2001 Artists Rights Society (ARS), New York / VG Bild-Kunst, Bonn. (B)
Analysis of Paul Klee’s With Green Stockings. The patches of tone correspond to
patches of color in the painting (yellow for the arms and green for the stockings).
These patches highlight line segments based on a denotation system in which lines
denote volumes. Numbers 1 through 6 mark anomalous changes in meaning along
the lines: for example, 1 indicates the change from a system in which the line
denotes a contour to a system in which the line denotes a volume. The single aster-
isk marks a false attachment giving rise to an anomalous T-junction. The double
asterisks mark false attachments between features in the picture and the picture
frame: in this case, the edge of the blotting paper. The arrows indicate anomalous
end-junctions. See plate 3 for color version.
141 Optical Laws or Symbolic Rules?

The first, and perhaps more obvious way, is to think of the continuation
of this line past the edge of the skirt as an example of transparency: we can
still see the leg as it passes behind the skirt, as if the skirt were transpar-
ent. Alternatively, we can think of this anomalous junction as occurring
as the result of a false attachment (see the part marked “*” in figure 6.1B):
the line above the edge of the skirt represents the contour of a fold in
the surface of the skirt, but this contour is falsely attached to the line
of the leg. With Green Stockings also contains both normal and anoma-
lous end-junctions. If we follow the line of the contour of the skirt on the
right of the painting, up past the point at which it is caught in by the girl’s
belt, it ends in the normal way. On the left of the painting, however, a simi-
lar line simply ends in mid-air (see the “→” symbol). Similar anomalous
end-junctions occur in the contours of the ball above the girl’s head. In
terms of picture perception we can no longer interpret these end-junctions
as picture primitives denoting the points where contours end; instead, they
simply become marks on the surface of the picture.
Finally, like Gris’s Breakfast, Klee’s painting contains examples of false
attachments (see the part marked “**” in figure 6.1B) between features of
the picture and the edge of the frame: these occur where the lines repre-
senting the ends of the arms just touch the edge of the paper.
I have analyzed Klee’s With Green Stockings in some detail, and alluded,
though more briefly, to the pictorial devices in Gris’s Breakfast, because
these two paintings represent extreme test cases for the argument that the
spatial systems in pictures of this kind are “conventional” in the sense that
in them the relations between scenes and pictures are arbitrary and sym-
bolic rather than natural. At first glance both these paintings may sug-
gest that this argument is convincing. Clearly, neither of these paintings
can easily be accounted for in terms of the laws of optics, or by simply say-
ing that they reproduce the structure of the optic array, or the “light that
strikes the eye.” Nevertheless, I have tried to show that in neither case are
the relations between scenes and pictures arbitrary. Both paintings con-
tain numerous instances of pictorial anomalies, but nearly all these
instances can be accounted for by saying that they break or reverse rules
derived from optical laws.
My purpose in giving these examples is to show that it is sometimes
appropriate to describe the representational systems in terms of the laws
of optics, and at other times it may be better to describe them in terms of
pictorial rules. Psychologists and theoretical writers on theories of pic-
tures have tended to favor descriptions based on projective geometry and
142 John Willats

the laws of optics, perhaps because psychologists have concentrated more


on picture perception than on picture production. Artists, however, have
usually been more interested in the rules involved in picture production,
because these are the rules they have to use in their paintings and draw-
ings. Some of these rules, like the perspectival rules about the orthogonals
converging to a vanishing point, have been used quite consciously, but it
seems probable that many of them operate at an unconscious level, like
the rules of language. However, one method of describing the spatial sys-
tems in pictures does not preclude the other: both methods are necessary,
and this duality seems to be an inherent property of representational
systems. Deciding which is the better way to describe these spatial sys-
tems ought to depend on the outcome of detailed investigations into the
spatial relations between scenes and pictures, rather than on some a priori
theory about the general nature of pictorial representation. As Stephen
Hawking has pointed out, there were for at least two hundred years two
hotly debated theories about light: “One, which Newton favored, was that
it was composed of particles; the other that it was made of waves. We now
know that both theories are correct . . . [and that] for some purposes it is
helpful to think of particles as waves and for other purposes it is better
to think of waves as particles” (Hawking 1993, pp. 81 and 56). I suggest
that something similar seems to be true of the representational systems in
pictures.

ACKNOWLEDGMENT

I am grateful to the University of Birmingham, Department of Psychol-


ogy, for financial help in enabling me to attend the Conference “Recon-
ceiving Pictorial Space” held at the Zentrum für interdisziplinäre
Forschung,” Universität Bielefeld, at which a shorter version of this
chapter was presented.

Notes
1. For example, Hagen regards all the representational systems that can be defined
in terms of projective geometry as perceptually equivalent, describing them as
varieties of “natural perspective” but argues that: “The selection from among the
representational options of natural perspective in terms of bias, preference, value
or function is determined by cultural convention” (1986, p. 83).
2. In practice, of course, photography nearly always involves some human
intervention.
3. Illustrated in Kemp (1990, fig. 390).
143 Optical Laws or Symbolic Rules?

4. Kemp (1990, fig. 278, pp. 145, 146). See also Kemp’s account of the working
methods of Pieter Saenredam, on pp. 114–116, and see also Willats (1997), on
pp. 190–193.
5. Illustrated in Willats (1997, fig. 3.2).
6. Illustrated in Dubery and Willats (1983, fig. 57, p. 54).
7. Children’s early drawings are usually composed of marks consisting of lines,
but it is crucial to distinguish between the marks in a picture, and the more
abstract notion of the picture primitives that these marks represent (Willats 1992a).
8. Illustrated in Gablik (1985, fig. 196).
9. Illustrated and analyzed in Willats (1997, pl. 6, pp. 275–279).
10. Illustrated and analyzed in Willats (1997, pl. 5, pp. 264–267).
11. Illustrated and analyzed in Willats (1997, fig. 12.3, pp. 280–281).
This page intentionally left blank
Chapter 7
Perspective, Convention, and Compromise

Robert Hopkins

7.1 THE ISSUE

What is special about picturing in perspective? The question needs clari-


fying. In one sense, all picturing is perspectival, for there can be no depic-
tion that is not from some point, or points, of view. In another sense, only
some pictures are perspectival, or in perspective. In this sense, perspecti-
val pictures are those conforming to the rules of some set of drawing sys-
tems. This is the sense relevant here, but it is hard to clarify satisfactorily.
One thing is clear: the required conformity is not historical. It matters not
how the picture was actually produced but whether the finished product
has the features the rules prescribe. The unclarity lies in just which draw-
ing systems are the relevant ones, and what, if anything, unites them as a
group. I will try to proceed without addressing these issues, returning to
them only at the close of the chapter.
Unfortunately, even granting this omission does not leave our question
clear. What sort of “specialness” is at stake? Enthusiasts and skeptics
about perspective sometimes frame the issue between them as whether or
not perspectival picturing has a special “authority” or “unique validity”
(Hyman 1992), but what does this mean? We can make some progress by
appeal to the idea of convention, provided we give that notion a clear con-
tent. As I will use the term, for a practice to be conventional is for it to be
one of two or more equally good solutions to a problem, all of which are
available to a given community, which is adopted by each member of that
community simply because it is adopted by the others, it being common
knowledge that this is so.1 In order to apply this notion to the matter in
hand, we need to specify the problem that picturing in perspective solves.
I will make a suggestion on this score at the chapter’s end. Given some
such specification of the problem, the question of perspective’s status as
146 Robert Hopkins

conventional or otherwise will center on whether its solution to that prob-


lem is uniquely advantageous.
Is the question, then, whether perspectival picturing is conventional,
in the sense defined? Not quite. For the skeptical, conventionalist answer
might take a more or a less radical form. We have as yet said nothing
about picturing in general. Given a specification of the problem it solves,
we can ask of it too whether it is conventional. The two skeptical positions
on perspective both hold that picturing in general is a matter of conven-
tion. The more radical view is that even given the conventions constitut-
ing picturing in general, the further practices definitive of perspectival
picturing are themselves a matter of convention. The less radical posi-
tion is that given those general picturing conventions, perspective is not
conventional. That is, given that more general conventional practice, per-
spectival systems solve their specific problem better than rival solutions.
An analogy may help. It is a conventional matter which side of the road
we drive on. But, given that practice, it is not a matter of convention
which direction one ought to look before stepping out from the curb. If
the problem is that of avoiding being run over, the background conven-
tions render one solution better than the alternative and hence not itself
conventional. The less radical skepticism about perspective sees it as play-
ing a precisely parallel role. We need to phrase our question so as to leave
room for this sort of view.2
We can thus formulate the question as follows: Is perspectival picturing
(a) a matter of convention; (b) a nonconventional consequence of those
practices, themselves conventional, which constitute picturing in general;
or (c) neither of the above? As noted earlier, this question is incomplete
without specifications of the problems picturing in general and perspec-
tival picturing in particular are, respectively, intended to solve, and with-
out some definition of the different perspectival drawing systems. I think,
however, that the question is sufficiently clear for discussion to proceed.

7.2 A THEORY OF PICTURES

I will answer the question by appeal to a particular account of pictorial


representation. That account needs more exposition and defense than I
can give here (see Hopkins 1998). If it is of interest anyway, that is because
it points the way to a compromise between the enthusiastic and skeptical
positions on perspective. Even if the account is rejected, perhaps some
of its rivals can adopt the structure of that compromise. The compro-
mise emerges in sections 7.5 and 7.6. Section 7.5 shows how the view can
147 Perspective, Convention, and Compromise

accommodate many of the ideas that suggest perspective is a matter of


convention. Despite that, section 7.6 explains just what is, on my account,
special about perspectival picturing, in terms that are, in one respect at
least, as strong as any enthusiast about perspective could want. But to
see how this combination of responses emerges, we need first to expound
the view. This section does that, and the next two elaborate and reject an
objection, one particularly salient here. This objection is that the view
seems unable to acknowledge the existence of any picturing other than
that in perspective.
The account begins from Richard Wollheim’s thought (1987, chap. 2 B)
that what is distinctive about pictorial representation is that it engenders
a special kind of visual experience. I follow Wollheim in calling this expe-
rience “seeing-in.” It is a necessary condition on a surface depicting some-
thing that that thing be seen in the surface. But the condition is only
necessary. What is further required is that something makes it right to
see that thing there. The requisite “standard of correctness” is supplied
either, in the case of nonphotographic pictures, by the intentions of the
artist, or, in that of photographs, by some special causal relation between
the marked surface and the thing visible in it.
Where I differ from Wollheim is in my account of seeing-in itself. In my
view, seeing-in is a form of experienced resemblance. To see something in
a marked surface is to experience the marks as resembling that thing in
a certain respect. The immediate problem is to say what this respect is,
because we might think that the differences between picture and object
are at least as obvious as any similarities. It’s tempting to think that the
picture is experienced as resembling its object in respect to some shape
property. But, given the manifest differences in 3-D shape between (most)
flat canvases and what they depict, what can this property be? There is an
answer to this question. It finds its neatest expression in the work of the
eighteenth-century philosopher Thomas Reid.
Reid distinguishes between two properties, which he dubs “real” figure
and “visible” figure (1764, chap. 6, sec. 7). The former is just 3-D shape.
The latter is in effect the shape things have if we ignore the dimension of
depth. It is a matter of the directions of various parts of the object from
the point of view. Thus, if I see a long table stretching away from me, dif-
ferent parts of the table lie in different directions from my point of view.
For instance, the two distant corners lie in directions different from each
other and different from the two directions in which the nearer corners lie.
The far right-hand corner lies to the right of the far left-hand corner, and
above the near right-hand corner. And the directions in which the two
148 Robert Hopkins

distant corners lie are closer to each other than are the directions of the
two nearer ones.3
I make two observations about visible figure. The first is that it is a gen-
uine property of things in our environment. For the point of view is just
the actual location from which the world is seen, and visible figure is sim-
ply a matter of the spatial relations in which parts of the object stand to
that location. This is a complex property determined by the object’s 3-D
shape and orientation to the point of view, but it is distinct from either of
these. Reid realized this, despite the misleading contrast with “real” figure,
for he described visible figure itself as “a real and external object to the
eye” (ibid., chap. 4, sec. 8).
The second observation is that visible figure deserves its name: it is
something we see.4 My comments about the directions of corners of the
table do not merely describe how the world is laid out. They also capture
how my experience presents that arrangement. I see the directions in
which the various corners lie, and see, for instance, that opposite points
on the table’s long edges lie in directions ever closer together, the farther
away from me they are. Of course, as with any property we perceive, our
experience can be misleading and will always be imprecise in certain
respects. When a half-submerged stick looks bent we misperceive, not just
its 3-D shape, but its visible figure too. And no visible figure is perceived
with complete precision: there is always a degree of specificity beyond
which experience neither represents the object as having this visible figure
or as having that one. Indeed, imprecision is an important feature of our
perception of this property. For experienced visible figure is only as deter-
minate as the point of view it involves, and in general our visual experi-
ence presents the world not from a point but from a zone in environmental
space large enough, in binocular vision, to include the actual location of
both eyes.5 But the observation stands: visible figure is seen, albeit with
varying accuracy and precision. This is just as well. Visible figure can
hardly be central to making and understanding pictures, those represen-
tations so closely bound to vision, unless visible figure is something vision
makes available.
Reid noted that it is the artist’s job “to hunt this fugitive form [i.e., vis-
ible figure], and to take a copy of it” (ibid., chap. 6, sec. 7). I more or less
agree. For it is possible for a picture to resemble its object in visible figure,
even though they differ considerably in 3-D shape. A picture of our table,
for instance, might represent the four corners of its top by marks which lie
in just the same directions from my point of view as did the corners them-
149 Perspective, Convention, and Compromise

selves, when I stood before the table. The marks lie in the same directions,
from the relevant point of view, as did the corners, even though, by lying
at different distances in those directions, they form a configuration very
different in 3-D shape from the table itself. But Reid’s formulation is not
quite right. What is crucial for pictorial representation is not actual resem-
blance in visible figure but that the marks be experienced as resembling the
depicted object in this respect. To experience this resemblance is to see the
object in the picture. And provided some intention or causal connection
renders it right to experience the picture in that way, it is the picture’s sus-
taining such experiences of resemblance in visible figure that constitutes its
depicting that thing. Thus we have the basics of an account of picturing.
What are the implications of this view for our question about perspective?

7.3 A CHALLENGE: ACCOMMODATING THE VARIETY OF PICTURES

This issue is best approached through a challenge to the resemblance view.


Surely at least some perspectival systems are designed precisely to preserve
the visible figure of depicted objects. Consider Alberti’s claim that when
painters cover a plane with colors, “they should only seek to present the
form of things seen on this plane as if it were of transparent glass. Thus
the visual pyramid could pass through it, placed at a definite distance with
definite lights and a definite position of centre in space and in a definite
place with respect to the observer” (1436/1966, p. 51).
Isn’t this strikingly reminiscent of my description of how a picture of
the table might preserve its visible figure? The only difference seems to be
that talk of the visual pyramid substitutes for that of directions from the
point of view. Now, whatever Alberti was doing, he was attempting to
prescribe how pictures should be fashioned. His injunctions presuppose
that not all pictorial representations are this way; otherwise there would
be no point urging that they should be. But if pictures fitting his prescrip-
tions preserve visible figure, presumably those that don’t fit them will not
do so. And if these other pictures don’t preserve visible figure, how can
they meet the resemblance view’s main condition on their counting as pic-
torial representations, namely, that they be experienced as preserving it?
The view purports to be an account of picturing per se, but it seems at
best to cover a tiny fragment of the pictorial realm, picturing in Albertian
perspective.
Of course, there are other perspectival drawing systems. Perhaps they,
or at least some of them, succeed in preserving visible figure too. But this
150 Robert Hopkins

at best goes only a small way toward meeting the challenge posed. For
even if all picturing in perspective meets the view’s criterion for picto-
rial representation, that still leaves all other picturing, quite implausibly,
excluded.6 In effect, the resemblance view seems to have taken what is true
of a subset of picturing, a subset at best as large as that of all picturing in
perspective, and made it definitive of picturing per se. And this leaves it
giving implausible answers to the two questions before us. To the question
“what is pictorial representation?” it gives an implausibly narrow answer,
in effect one applying only to perspectival picturing. If so, it also gives the
wrong answer to the question “what is special about picturing in perspec-
tive?” For it implies that it is special precisely in being the only picturing
there is.
To save the resemblance view, we need first to argue that it covers all
the rich variety in ways of picturing. But, having done that, we need also
to find some other account of what is special about pictures in perspective.
The next section undertakes the first of these tasks. Because its main aim
is negative—that the view can evade the current objection—readers more
interested in my positive assertions about picturing and perspective may
choose to skip this section.

7.4 MEETING THE CHALLENGE

Let us begin by trying to sharpen the challenge. Consider a particular


object, say a long table, at a particular distance and orientation. The chal-
lenge concerns the range of things the resemblance view can allow to
count as pictures of that thing, at that distance and orientation. Only a
limited range of marks will preserve the visible figure of the table when
so positioned. They might differ in color, texture, and sharpness of line,
say, but not in their fundamental positions on the surface. Now, it is true
that the view does not analyze depiction in terms of actual resemblance in
visible figure—experienced resemblance is what matters. And it is quite
possible that picture makers have learned to elicit such experiences even
when the resemblances themselves are missing. After all, they have done
just that for colors, using context to ensure that a pigment is seen as the
color of spring foliage, for all that the two are really rather different in
hue. But there are strict limits on such trickery: a pink pigment cannot be
seen as resembling one that is lime green, whatever the context. And this
is because, generally, our perception of features of our environment is,
though not perfect, mostly accurate. If this is true of resemblance in visi-
ble figure, there will be relatively little slippage between those marks actu-
151 Perspective, Convention, and Compromise

ally preserving visible figure and those experienced as doing so. More or
less, then, only those marks preserving visible figure will count as depict-
ing the table. The challenge to the view is to make room for a wider range
of marks, counting as depictions of the table.
Showing that the resemblance view can meet this challenge is a complex
matter. Here I can do no more than sketch some of the ingredients in a full
reply. In this section I make three points. All three embody the basic
observation that marks depicting the table need not match any further in
content than that.
The first point is that not all marks depicting the table need depict it
accurately. Some, that is, might depict the table as having properties it
does not actually enjoy. It is obvious that, in general, pictorial misrepre-
sentation is possible. A caricature of Tony Blair need not show his nose as
being the shape it is and may indeed distort his features quite drastically,
while still depicting him. So there is no reason for every picture that
depicts our table to show it as the shape or color it really is or as having
the appropriate number of legs. Such misrepresentations differ in content
from each other and from accurate depictions (save, of course, that at
least one element is common to the content of all: the table). Now, with
respect to such groups of pictures, the resemblance view, far from pre-
cluding differences in the marks composing them, actually expects such.
Like any account of depiction, it seeks to say what it is for a picture to
have its pictorial content. For instance, it will say that for something to be
a picture of the table as having six legs, a round top, and so forth, is for it
to be experienced as resembling in visible figure the table with six legs, a
round top, and so on; but that to depict the table as round topped, and so
forth, but with seven legs, the marks must be experienced as resembling
the table with seven legs, a round top, and so on. The view does not say
how the surfaces must be marked to sustain these experiences. But given
the rather limited slippage between actual resemblance and experience of
it, it will at least be comfortable with the idea that different experiences of
resemblance will be sustained by differently marked surfaces. Here then is
one simple way in which marks depicting the table, from the specified
angle, may differ from one another, so that not all of them preserve the
table’s actual visible figure.
The thoughts underlying this move are central to the second point,
too, so it is worth spelling them out. Pictures need not be of particular
things: a painting can depict a horse without their being any answer to the
question of which horse it represents. But let us for the moment stick to
pictures that do depict particulars. In effect, I have said that for these
152 Robert Hopkins

pictures the view makes central, not experienced resemblance to the pic-
ture’s object as it actually is, but experienced resemblance to that item
with whatever properties it is depicted as having. As I have noted, it is in
a way obvious that this is how things should be. For only thus can the
view perform its main job, of stating what it is for a picture to have a given
pictorial content. For because the content of a picture is always richer
than just the representation of some particular, and because there is no
need for the properties filling out that content to be limited to ones the
particular actually enjoys, the view must say that what matters is that
resemblance to the object be experienced with whatever properties the
picture ascribes.
However, despite this obviousness, discussions of resemblance views
have always overlooked this point. There are several reasons for this.
Some are good reasons, although there is no space here to argue that they
do not justify rejecting the claims of the last paragraph.7 But others are
very poor. Among the poor reasons is a fixation on the idea of perspecti-
val systems as ways of projecting particular objects onto plane surfaces.
Thinking that this is what perspectival systems do, and thereby thinking
of nonperspectival picturing as another way, or set of ways, of achieving
the same result, makes it especially hard to grasp the thought that the test
of the resulting marks is experienced resemblance to whatever the picture
depicts and not experienced resemblance to the original projected object.
Yet in fact the notion of a perspectival system, at least in its full gener-
ality, has nothing to do with the projection of real things. As I noted at
the start, rather than conceiving of perspective in this historical way, it is
better to think of it as a set of conditions that completed pictures may or
may not meet. So conceived, perspective can apply, or fail to apply, to pic-
tures whether they are of particular objects represented accurately, of par-
ticular objects with properties they do not in fact enjoy, or not of any
particular object at all. If we do not conceive of perspectival picturing in
this way, it will seem less interesting than it really is. For, as noted, pic-
tures in general are not limited to representing particulars nor to rep-
resenting any particulars they do depict as having the properties they
actually do. So picturing in perspective, if it is conceived as a way of
projecting particular objects onto surfaces, does not, whatever its other
benefits and limitations, offer the same range of possible contents, in terms
of fundamental logical kinds, as picturing in general does.
This first point begins the process of broadening the range of pictures
the view can countenance. But it only gets us so far. For misrepresenting
pictures too might either be in perspective or not. We have seen how very
153 Perspective, Convention, and Compromise

different marks might all depict our table, by ascribing different properties
to it. But nothing said so far shows how those pictures might be other than
perspectival. So let me turn to the second point. This is to identify a sec-
ond way in which pictures of the table might differ in content: rather than
being inaccurate, their content might be imprecise. One might represent
the table’s shape very precisely, while another represents it as merely
roughly round. For pictorial content to be imprecise is for there to be no
answer, beyond a certain level of detail, as to which of two distinct prop-
erties the picture ascribes to its object.8
It is clear that pictures can have imprecise content. A quick sketch
might represent nothing more than a figure’s rough shape and posture. It
should be equally clear that the resemblance view can allow for this fea-
ture of depiction. As we saw, what matters is experienced resemblance to
the object as depicted. There is no reason why marks should not be expe-
rienced as resembling something with fairly indeterminate properties. But
it will help meet the challenge to spell out just what such an experience
does and does not involve.
What is not required is that the subject misperceive the marks, or only
perceive them with some rather limited level of precision. I can see a
child’s drawing as resembling in visible figure our table, with the prop-
erties the picture ascribes, even if I see the wobbly lines that make up the
drawing with perfect clarity. Nor is it required that I experience those
marks as only resembling the (relatively indeterminately characterized)
table to a certain degree. In the preceding text I made no mention of
degrees of resemblance in explaining how the view accommodates mis-
representation, because there is no need to appeal to that notion. For all
I said was that we experience pictures as resembling their objects, as
depicted, perfectly, even when the objects as they actually are do not share
the marks’ visible figure. And the same holds here. I may experience the
child’s wobbly lines as perfectly resembling a table in visible figure, pro-
vided the table in question is only roughly characterized, for example, as
more or less cuboid, rather longer than it is wide, and so forth. What is
required for such experiences is that although I may see the detailed fea-
tures of the marks, only certain aspects of them are salient in my experi-
ence of resemblance. It is the rough shape of the line, not its meandering
edges, that is prominent for me in seeing the resemblance to the table,
indeterminately characterized.
How exactly do these observations about imprecise pictorial content
help the resemblance view respond to the challenge? Different pictures of
our table might, without misrepresenting it, differ in content from the
154 Robert Hopkins

perspectival representation, by depicting less determinate properties. Again,


where there is a difference in content our view will predict, rather than pre-
clude, that there at least may be a difference in the marks having that con-
tent. For, again, if two sets of marks are to be experienced as resembling
different things, one way to achieve this is for them to differ from one
another.9 One might think that these nonperspectival pictures succeed in
depicting the table only insofar as they approximate to the perspectival
depiction. But that is quite wrong. They are experienced as resembling
a table somewhat indeterminately characterized, but there is no sense in
which the actual table, with its real properties, provides some privileged
way of filling in the indeterminacies. If one of them depicts the table as
roughly cuboid, it matters not what the precise shape of the real table
is. That is no more relevant to our characterization of the experience of
resemblance than any other way of filling in the gaps. So there is no sense
in which the perspectival picture preserves perfectly the visible figure
that the other pictures merely roughly preserve. And, as noted, there is no
room for the thought that the nonperspectival pictures are experienced as
resembling their objects to some weaker degree than the picture in per-
spective. So the claim of nonperspectival pictures with relatively indeter-
minate content to depict the table is not in any way parasitic on, or less
secure than, the claim of the perspectival picture to do so. Also, given that
in different pictures with imprecise content different aspects of the marks
composing them may be salient, it seems that appeal to such imprecision
widens considerably the range of marks that count as depictions of the
table.
The third point develops the theme of imprecision. Thus far I have
assumed that where pictorial content is imprecise, so will be the experience
determining that content. That is, if pictures ascribe relatively indetermi-
nate properties to their objects, that is because they are seen as resembling
objects with indeterminate properties. But there is another means by
which pictorial content can be imprecise. Remember that depiction is not
merely a matter of experienced resemblance to some item. It also requires
that some standard of correctness renders it right to see the marks as
resembling that thing (section 7.2). It is possible for these two elements in
the view to come apart. In general, this merely prevents depiction from
occurring, as when patterns in the frost on a window pane fail, despite
being seen as resembling a figure of some kind, to depict that figure, or
anything else, because nothing makes it right to see them as resembling
something (Wollheim 1987). But in certain cases a mismatch between the
155 Perspective, Convention, and Compromise

experience of resemblance and the standard of correctness can merely


serve to modify pictorial content. This occurs when what the standard
makes it right to see the picture as resembling is less determinate than
what the picture is seen to resemble. An example is provided by stick
figure pictures. We see these as resembling very thin people, with swollen
heads. But, although they depict people, they do not depict them as oddly
shaped. For, although the drawer’s intentions make it right to see the
resemblance to people, it is not intended that resemblance to emaciated
figures be experienced. The drawings depict people, but it is simply inde-
terminate, within certain broad limits, what shape they are. Here, then,
pictorial content is imprecise in ways in which the content of the experi-
ence of resemblance is not.
A relevant instance of indeterminacy attained by these means may be
pictures in oriental perspective. These preserve their objects’ visible figure
but only that visible figure formed at a point very distant from, and stan-
dardly some distance above, the depicted scene (Hagen 1986). Thus there
is no trouble understanding how these pictures can be seen as resembling
their objects in visible figure, despite the fact that they do not deploy clas-
sical Western perspective. It seems to me an open question whether they
depict their scenes from the high, distant, viewpoint described. But even if
they do not, the resemblance view can easily accommodate this difference.
The content of the picture is less precise, with respect to viewpoint, than
is what is seen in it, because what is seen in these marks and what it is right
to see therein comes apart in the way I have described.
There is more to be said about this alternate method for generating
imprecise content. In particular, we need to know how those who see
the pictures work out just what is depicted, that is, which aspects of
their experience to rely upon in interpreting the pictures and which to set
aside. I will not go into these matters here (but see Hopkins 1998, chap.
6). My point now is that provided these questions can be answered, this
second route to indeterminacy further widens the tolerance of the resem-
blance view. As the case of stick figures illustrates, marks very different
from those required by perspectival systems can depict, given this latest
refinement to the view.
I said at the start of this section that my three points would all exploit
the possibility that different pictures of our table might differ in content.
This is true, but the second and third in particular also allow us to see
how rather different marks might nonetheless exactly match in picto-
rial content. They can do so if only certain aspects of them are salient in
156 Robert Hopkins

determining what the marks are seen as resembling. Or they can do so by


supporting experiences of resemblance to things with different properties,
provided only the common aspects of those resembled items are under-
written by standards of correctness and thus retained in the content of
the pictures.

7.5 COMPROMISING WITH THE CONVENTIONALIST

The points of the last section go a good way toward meeting the challenge
to the resemblance view. Marks very different from those preserving our
table’s visible figure might nonetheless depict it, from the right angle, pro-
vided they either misrepresent it or represent its properties only impre-
cisely, or, of course, do both. But I promised to do more than merely
deflect objections. I promised some form of compromise with those who
take depiction itself, and perspectival picturing in particular, to be either
properly conventional or a nonconventional consequence of conventions.
How does the view yield that compromise?
We can see room for compromise by getting clear about the view’s
ambitions. It seeks to analyze depiction in terms of the experience to
which it gives rise, seeing-in. To do that, it needs to tell us what that expe-
rience is. It is the experience of resemblance in visible figure. But charac-
terizing an experience is not the same thing as describing its causes. So far
our view has not said much about how the world must be for a given sur-
face to bring a given spectator to see a given thing in it. It is not entirely
without implications for this question. If seeing something in a set of
marks is a matter of experiencing them as resembling that thing in visible
figure, then, given some basic facts of optics and psychology, the only
marks that will do, except in exceptional circumstances, are ones at least
approximating to the object’s visible figure. And although the last section
argued that there is no need for the approximation to be very exact, it does
not follow that it need not hold at all.10 But the causal condition thus
stated is at best necessary. Many other factors will come into play in dic-
tating whether a given experience of resemblance in fact occurs. Saying
what these factors are is no part of the brief of our position. Its goal is to
analyze seeing-in, and thereby pictorial representation, not to describe the
empirical facts about when that experience is engendered. However, the
view can at least acknowledge these facts. It is here that the seeds of com-
promise can be found. For the view can allow that many of the factors
empirically determining a given experience of resemblance are precisely
157 Perspective, Convention, and Compromise

those which the skeptic about perspective wished to emphasize. I offer


three examples.
The first factor is the nature of the spectator’s general perceptual envi-
ronment. We can see that this environment must play a role in determin-
ing what folk see in surfaces, by better understanding why that cannot be
determined by preservation of visible figure alone. The problem is that a
given visible figure is compatible with an infinite range of 3-D shapes.11
The attraction of placing visible figure at the core of an account of pic-
turing is precisely that objects of different shape, and in particular pic-
tures and what they depict, might share visible figure. They need only be
arranged differently, that is, at different distances, along the same set of
directions from a given point. But by the same token any other arrange-
ment differing from these two only in respect of depth will likewise share
that visible figure. So the fact that a picture preserves the visible figure of
a particular 3-D shape cannot alone determine what is seen in it. For we
might equally see therein one of the many other shapes the visible figure
of which is thus preserved.
What, then, does determine which of the infinitely many compatible
3-D shapes we experience a given set of marks as resembling in visible
figure? In part, the answer is surely our general perceptual environment.
We see cubes in pictures of them, rather than irregular trapezohedrons,
because our world contains far more of the former than the latter. Such
examples may seem to offer little comfort to the skeptical, conventional-
ist, cast of mind. For these folk want to stress the variation across cultures
in the reception and use of perspective; whereas many of the environmen-
tal features whose role we are now describing will be common to all cul-
tures. But the structure of the issue here is more general. A lot of people
look more or less alike, some very much so. What determines that people
see some of these look-alikes rather than others in their portraits and
holiday snaps? It is not resemblance in visible figure, because what so
resembles one look-alike will equally resemble all. So the answer must lie
in past perceptual experience: a person has seen one look-alike but had no
contact with the other. Here there is variation across different groups of
people of just the sort the skeptic seeks to highlight. The resemblance view
can acknowledge the role of that variation not by offering a convention-
alist account of depiction but by allowing the variation a role in the causal
determinants of seeing-in.
The second factor I want to discuss is closer to the skeptic’s heart. This
is the spectator’s pictorial education. It is a matter of the sorts of pictures
158 Robert Hopkins

to which a person has been exposed and the ways in which those have
affected that individual’s perceptual sensitivities. Although few have real-
ized this, the resemblance view is easily able to acknowledge the role such
factors play. As Reid noted, visible figure is “fugitive,” that is, it is hard to
perceive with precision and accuracy. The less sharp one’s perceptual
grasp of it, the less sensitive one will be to a picture’s failure to preserve it.
In consequence, many nonperspectival pictures may well have been expe-
rienced as resembling their objects in visible figure perfectly well, for all
that they don’t in fact resemble them more than approximately. For they
may have been intended for, and enjoyed by, viewers whose perception of
visible figure was less than ideally acute. And one result of the rediscovery
of perspective may have been to sharpen viewers’ ability to perceive visi-
ble figure, thereby altering which marks will be experienced as resembling
things in that respect.
Here the compromise with the skeptic is considerable. Our view allows
that picturing essentially involves something independent of conventions,
the generation of a particular experience, but the view also allows that cer-
tain pictures may themselves alter which pictures are experienced in that
way. The suggestion is that perspectival pictures may be one such case,
altering viewers’ general perceptual propensities in order to bolster their
own position as the marked surfaces most likely to elicit the required expe-
riential response. The point can be illustrated by tackling an argument of
Nelson Goodman’s, one intended precisely as an objection to perspec-
tive’s special status.
One of Goodman’s points against perspective was, in effect, that not
even pictures in Albertian perspective preserve visible figure (1969, chap.
1). The Albertian system requires receding horizontal parallels to be rep-
resented by converging lines but not so parallels receding in the vertical
plane. As Goodman observed, this is not true to the geometry. The differ-
ence between the directions of opposing points on the parallels narrows
just as rapidly with recession in the vertical plane as in the horizontal.
What Goodman overlooked, however, is that this discrepancy may reflect
an aspect of our experience of visible figure. Setting aside pictures for a
moment, consider our face-to-face experience with things. Suppose that in
such experience we are more sensitive to this shift in the direction of points
on receding parallels when the recession is horizontal, as along railway
tracks, than when it is vertical, as with the sides of a building viewed
from the street. Then we would expect pictures that are experienced as
preserving visible figure to exhibit precisely the discrepancy noted. Sys-
159 Perspective, Convention, and Compromise

tematic discrepancies in our face-to-face experience of visible figure will


dictate analogous discrepancies in our depictive practices. This provides a
response to Goodman’s objection, but it also suggests another point more
relevant to present purposes. For it may be that our differing sensitivities
to vertical and horizontal recession is as much effect as cause of the
Albertian system. Perhaps the prevalence of picturing in Albertian per-
spective has sharpened our perception of one of these aspects of visible
figure over the other.
Of course, it is only a supposition that there are the systematic discrep-
ancies in experience just described. And it is only a hypothesis that any
such discrepancies might be fed by, as much as support, the pictorial sys-
tems we have adopted. These claims need testing empirically. But they can
serve my purposes prior to such investigation. For the supposition alone
is enough to neutralize Goodman’s objection to perspective’s privileged
status, pending further investigation. And the hypothesis serves to illus-
trate that to the extent that the skeptic is right to think that accultura-
tion affects which pictures look natural or right to us, the truth in such
claims is consistent with the resemblance view. For the hypothesis, pro-
duced within the framework of that view, is that a perspectival system has
altered our perceptions in such a way as to ensure its own entrenchment
as a way of generating pictures that look right. Suspicion that perspec-
tive’s success is down to such bootstrapping is surely one of the motors of
skepticism about it.
The third factor determining experiences of resemblance is another one
that might vary across cultures: the way we view pictures. Once again, we
can approach the point via a potential objection to perspectival pictur-
ing’s special status.
In Raphael’s The School of Athens, 1510/1511, a figure on the far right
holds a sphere, resting on the fingertips of his upturned hand. The sphere
is depicted by a circular mark. However, classical perspectival systems
suggest that thanks to its oblique position away from the vanishing point
of the picture, the sphere should be represented by an elliptical patch of
paint. This preserves the visible figure a sphere would have at the single,
highly precise, point in depicted space from which the entire scene might
be viewed and from which Raphael depicts the scene. In the nineteenth
century, Jules de La Gournerie tried substituting, on an engraving of
The School, an ellipse for Raphael’s circle. However, the effect looked
quite wrong, whereas Raphael’s circle looks exactly right. The reason is,
once again, that actual resemblance in visible figure is no guarantee of
160 Robert Hopkins

experienced resemblance therein. What determines the discrepancy here?


The answer is that the way we view pictures affects which marks we see as
resembling others in visible figure. Were we to view the Raphael from the
point, in front of the marked surface, at which it exactly matches the vis-
ible figure, at the pictorial point of view, of the depicted scene, and were
we to view it from there with one eye, then the elliptical patch would look
right. That is, it would be seen as resembling an off-center sphere in visi-
ble figure. But we don’t normally look at pictures in that way. We move
about, positioning ourselves at different distances from the marked sur-
face and at different points along its length. We use both eyes. We let our
impressions build over time. All these factors dictate that the circular
patch is the one we see as resembling a spherical, off-center ball; and thus
that such a patch is what depicts the ball. Things would be different were
we the sort of viewers, with different habits of attending to pictures, La
Gournerie’s unduly rigid application of perspective supposed us to be.
Then an elliptical patch would carry the content of our circular one.12

7.6 WHAT IS SPECIAL ABOUT PERSPECTIVE?

The last section tried to show that the resemblance view can accommodate
some of the factors that have impressed those skeptical about perspec-
tive’s special status. Section 7.4 argued that the view is in no way obliged
to restrict depiction to picturing in perspective. This leaves one issue out-
standing. We need to say what is special about perspectival picturing,
according to the resemblance view. If it’s not the only picturing there is,
and if the truth about both depiction in general and perspectival depiction
in particular leaves room for at least some of the culturally relative fac-
tors the skeptic emphasizes, then how should we answer the question with
which we began this chapter? Is perspectival picturing conventional or a
consequence of conventions? And if it is not conventional, what is the
problem that it solves better than any rivals?
Perspectival picturing is not a consequence of conventions, for pictur-
ing itself is not conventional. Pictorial representation essentially involves
the generation of a certain experience, one of resemblance in visible figure.
Generating this experience, in such a way as to effect communication, is
the problem that methods of picturing have to solve. More precisely, in
any particular case the problem is so to mark a surface such that it can be
seen as resembling, in visible figure, a certain thing or scene, with certain
properties, from a certain angle. Given our account of convention (section
161 Perspective, Convention, and Compromise

7.1), picturing will be conventional just if there are two or more equally
good solutions to each such problem, the choice between them being made
on the basis that others have also chosen likewise, it also being common
knowledge that this is what is going on. There are three reasons why this
condition is not met.
First, the range of possible marks in which the target content can be
seen is narrowly constrained. For despite all the concessions made above,
it remains the case that in general a causally necessary condition is that the
marks preserve, at least roughly, the relevant visible figure (section 7.5).
This excludes most of the possible ways of marking the surface from being
solutions to the problem in hand.
Now, those concessions were real. I allowed that other factors do play
a role in determining experiences of resemblance, factors other than the
marks’ actually reproducing the visible figure of the target scene. Nonethe-
less, none of these factors plays the role convention would require. For—
the second reason—they are not in general factors under the control of
a given group of picture makers or consumers. This is clearly true for
the general nature of the perceptual environment. We can control the
nature of what others or ourselves are in perceptual contact with but only
under special circumstances (e.g., persuading someone to take part in a
psychological experiment) and over very limited portions of space and
time. The limited context over which we thus exercise control plays a
highly restricted role in conditioning pictorial perception, compared with
the subject’s perceptual life history. The point also holds for pictorial
acculturation. No doubt for any pictorial education a viewer has had,
there is a different one that person might have undergone. But it does not
follow that either the viewer, or any picture maker currently attempting to
communicate with that viewer, could have chosen to give him the one
rather than the other. It is a matter of the history of the viewer’s culture
and own exposure, over many years, to various images. And something
similar is at least in part true of our habits of viewing pictures. One can,
by prescription or by constricting the viewer’s movement, in exceptional
circumstances limit the ways in which the picture is viewed. But even
here many of the key factors, such as the tendency to deploy eye or head
movement, reflect engrained perceptual habits and even basic facts about
visual physiology, which cannot be swept aside, however determined we
are to find a new way of doing things.
Perhaps these considerations do not exclude every relevant possibility.
Perhaps they leave some room at the margins, by manipulating the factors
162 Robert Hopkins

discussed, for alternative means to the same experiential end. But, and this
is my third reason that the aforementioned condition on conventionality
is not met, even if it is true that alternative means exist, it is not common
knowledge that it is. Different viewing habits, for instance, may affect what
people see a set of marks as resembling, but those people are not aware
that this is so. This will prevent them from knowing that what they are
responding to is one of two equally good solutions to a given pictorial
problem. And thus, insofar as the factors of section 7.5 meet the control
condition for conventionality, they do not meet the common knowledge
condition. Picturing, it seems, is not conventional.
Nor is it the case that perspectival picturing is conventional. It might
seem that my attempt to defuse the objection that the resemblance view is
hopelessly narrow forces just this result. For if perspectival picturing is
not the only sort there is, and if the problem solved by all pictorial systems
is just that of generating experiences of resemblance to certain things, it
may seem that there is nothing left for perspective systems to do better
than their rivals. If not, we would have in place at least part of what is
required for their adoption to be a matter of convention. But this line is
mistaken. There is something perspective systems do that others do not:
they are not the only pictures, but they constitute the only way to depict a
certain kind of thing. For only pictures in perspective depict detailed spa-
tial content. Or so at least I will argue. I cannot derive this conclusion
from the resemblance view alone. However, the only extra support need-
ed is a little theoretical reflection and some plausible psychological specu-
lation. Given these foundations, although nonperspectival pictures might
depict spatial detail, it seems that they do not in fact do so. Moreover, a
lot would have to change before they would.
The theoretical reflection is this. Consider properties that admit of
ordered variation. Colors are one example—there are many shades
between pure green and pure blue. Spatial properties are another—my
hand can occupy many positions between my two feet. One can represent
such properties, of course, in more or less precise ways—merely by speci-
fying rough hues or locations, or by specifying more determinate ones.
Now, some forms of representation allow one to represent instances of
these properties in a highy precise but piecemeal way. One could, for
instance, have a word for a particular shade of blue, without having words
for any but the roughest categories of the other colors. But depiction is not
a form of representation that allows for this possibility. If you can depict
an instance of some color, at a certain level of precision, your represen-
163 Perspective, Convention, and Compromise

tational resources must also allow you to depict other colors, with equal
precision. And likewise for spatial properties.13
Coupled with the resemblance view, this thought promises to ensure
that only pictures in perspective can depict precise spatial properties. For
suppose we consider a competing system—so-called inverse perspective.
In pictures drawn on these principles, receding parallel lines in pictorial
space are represented by lines on the canvas that diverge, in sharp con-
trast to perspectival systems, in which such lines converge. Thus, because
perspectival systems preserve visible figure, inverse perspective cannot do
so. Now, despite that, there is no reason to deny that inverse perspective
yields depictions. For an inverse perspective drawing of an oblong table
may indeed be seen as resembling, in visible figure, an oblong table. But
this is only obviously possible if the object the picture is experienced as
resembling is reasonably indeterminate—a table that is roughly oblong,
no more. So let us consider under what conditions, if any, such a system
would yield depictions of precise shape and other spatial features. They
must be conditions allowing for the depiction of any of a wide variety of
such shapes and arrays, or so our theoretical reflection suggests. So the
rules of inverse perspective will have to allow for projecting a wide variety
of such shapes and arrays onto surfaces. But if we are to appreciate the
detailed content of the pictures thus created, it seems we must be sensi-
tive to small differences between the way a given canvas is marked and the
ways in which it might have been. Now, those differences are precisely
between marks that do not preserve the visible figure of what they repre-
sent, and their not doing so is important to their representing, according
to the rules of inverse perspective, what they do. So an appropriate sen-
sitivity to the relevant details of the surface markings cannot, in all psy-
chological plausibility, accompany experiencing the marks as resembling,
in visible figure, something with the precise spatial features in question.
Thus, although the marks may depict something, and although they may
systematically represent spatial detail, they do not depict that detail. The
only conditions under which they would do so would be ones in which our
psychology were quite different from how it in fact is.
Of course, inverse perspective fails in a radical way to preserve visible
figure. However, any other way of marking surfaces that fails less spec-
tacularly will also confront the basic problem here. So the resemblance
view, coupled with our theoretical constraint on depiction and our sense
of what is psychologically possible for us, does seem to preclude any depic-
tion of spatial detail that is not depiction in perspective. If this conclusion
164 Robert Hopkins

seems outrageous, remember that it is restricted to a certain, highly spe-


cial, form of pictorial content. Once one focuses that thought clearly, I
think the conclusion loses its air of reductio and starts to seem plausible.
At least it allows us to say what is special about drawing in perspective.
The Renaissance did not invent depiction, but it did discover a way to
expand our depictive powers. For only with the articulation of perspec-
tival drawing systems did we acquire the ability to depict space with
precision.14
One issue remains. Which systems are the perspectival ones? I am not
sure what to say about this. If the speculations just offered are correct,
we could pick them out as those systems permitting depiction of spatial
detail. The cost of doing this, however, would be to trivialize my claim
that perspective is not a matter of convention. For I made good that claim
by saying that no rivals can solve the problem of depicting spatial detail,
an assertion the proposed characterization would empty of content. What
is needed, therefore, is a way of picking out the perspectival systems that
renders my speculation plausible without trivializing it. I am not sure
what, other than a list of particular systems, would do this. But I am also
not sure that this is particularly a problem for the position articulated
here. I suspect that everyone who wants to talk clearly about whether per-
spectival systems have a special status will have to define which systems
they mean, and to say, in independent terms, which problem those systems
are intended to solve. Although this chapter has not managed the first of
those tasks, it has at least made an attempt at the second.

Notes
1. This is a drastically simplified version of David Lewis’s much admired account
(1969).
2. Logic leaves room for another position, on which picturing is not conven-
tional but picturing in perspective is so. Although intriguing, this view is not
adopted by any writer known to me and is not discussed here.
3. Visible figure is what I have elsewhere (Hopkins 1998) called “outline shape.”
4. Do not be misled by Reid’s term. It is not part of the notion of visible figure
that it is the only spatial property, or “figure,” that vision represents. My view is
that others, including 3-D shape, are also seen. Although Reid disagreed (see
Hopkins, 2002), this claim is a further, and optional, element in his position.
5. This last is an empirical speculation but a plausible one. If right, it allows us to
dismiss an incipient worry. This is that because we see with two eyes, we cannot
see from a single point of view and hence that something is amiss in my claims
about the perception of visible figure.
165 Perspective, Convention, and Compromise

6. Given this, I need not at this point address the issue, set aside in section 4.1, of
which systems are the perspectival ones.
7. These reasons are (1) it is assumed the resemblance view is designed to do a
job it should not, in fact, attempt to do, namely, explaining what it is about the
picture that leads people to experience it as they do (see section 7.5); (2) it is cor-
rectly assumed that the resemblance view will want to give some role to how the
depicted item actually is, the appropriate role then being misunderstood; and (3)
the worry that if what matters is resemblance to the depicted object as depicted,
the view takes for granted part of what it is supposed to explain, thereby guaran-
teeing that its analysis cannot be sufficient. For discussion of these objections, see
Hopkins (1998).
8. As I will say, the content is imprecise, and it is indeterminate how things are in
the depicted realm.
9. This is only one way in which this might be achieved, because given the point
about salience, matching marks might, under the right conditions, differ in content
in virtue of different aspects of each being salient in the experience of resemblance.
Hence the view predicts, not that the perspectival and nonperspectival pictures will
differ but that they might.
10. In section 7.4 I said that the view makes no appeal to (experienced) degrees of
resemblance in characterizing seeing-in. Indeed, the notion of degrees of resem-
blance does not figure in any of the claims that constitute the view. It is quite con-
sistent with this that the view, coupled with some basic optics and physiology, has
implications involving (actual) degrees of resemblance.
11. Strictly, the relevant range is of different pairings of 3-D shape and orientation.
12. For La Gournerie’s experiment, see Pirenne (1970, p. 122). His engraving
seems certain to have been far smaller than Raphael’s original. This complicates
the relationship between the story and the moral I want to draw from it. However,
I think it obvious that the outcome would have been the same had La Gournerie
tampered with the Raphael itself, so I ignore this complication here.
13. This thought is really the offspring of Goodman’s account of depiction (1969).
It ignores some of his view’s difficulties by restricting itself to certain sorts of prop-
erties and by resisting the temptation to insist on infinitely precise representation.
But the most significant difference is that I do not offer it as a definition of depic-
tion, more as a test for it. For another use of a similar thought, see Lopes (1996).
14. John Hyman (1992) has reached somewhat similar conclusions independently.
It might be fairer to consider the Renaissance achievement one of rediscovery.
See, for instance, White (1987).
This page intentionally left blank
Chapter 8
Resemblance Reconceived

Klaus Sachs-Hombach

As the existence of different types and schemata of pictures demonstrates,


pictorial forms of representation are subject to historical and cultural
influences. This can be seen as supporting a “convention theory” of pic-
torial meaning but does not exclude a resemblance theory. This chap-
ter defends the latter view on the basis of an internalized resemblance
criterion.

8.1 INTRODUCTION

Imagine you are shown a passport photograph of an unknown person.


A short time later you recognize someone on the road as the person in
the photograph. This phenomenon of pictorial reference is common but
puzzling. Many interesting empirical findings, from, for instance, psycho-
logical and neurological research, cast light on it. In contrast to such
empirical approaches, philosophical analysis aims, first of all, to clarify
what is meant by “pictorial reference.” This requires us to give an account
of what pictures are, that is, to formulate a criterion allowing us to clas-
sify certain things as pictures, or, at least, certain signs as pictorial signs.
There are currently two fundamentally different lines of thought within
philosophical analysis that claim to be able to give such an account.1 One
holds that, like linguistic expressions, pictorial reference is convention-
ally determined (so that the nature of pictures cannot be defined by their
semantics); the other assumes a genuine relation between the picture and
its subject (so that the internal structure of the picture allows us to deter-
mine at least partially what the picture depicts). Consequently, theories of
the first group maintain that a picture could have looked quite different
while still fulfilling its purpose; a variety of forms is available to depict
something, and the choice depends on cultural influences. Theories of the
168 Klaus Sachs-Hombach

second group insist that a picture, in order to be considered to depict a


certain object or event, must have some essential properties in common
with the object or event depicted; so that at least all objects devoid of these
properties are eliminated as referents. A prominent theory from the first
group is Nelsm Goodman’s symbol theory.2 Resemblance theory belongs
to the second group.3
For some time now, the usefulness of resemblance theory has been
questioned. Certainly, the issue of resemblance criteria is highly prob-
lematic and requires further explication.4 It is Goodman’s merit to have
pointed out these difficulties. However, the conventionalist view (partic-
ularly when applied to our example) seems to me extremely implausible.
I would, therefore, rather defend the resemblance criterion, at least with
respect to clear cases of figurative pictures. The passport photograph is a
particularly strong example of this kind of picture, because it has intuitive
appeal. In section 8.2, I discuss the intuition addressed by this example in
more detail. In section 8.3, I present my own proposal, which attempts to
reconcile the two previously cited approaches. In my view, pictures are
perception-based signs.5 In a final section, section 8.4, I take a brief look
at the problem of pictorial reference and certain puzzling cases.

8.2 A FIRST INTUITION

The example of the passport photograph is meant to elicit a strong intu-


ition in favor of resemblance theory. With it, one points out that our use
of pictures is bound to a particular competence, which is different from the
competence one needs in verbal communication. We are, it seems, able to
understand the passport photograph immediately, without having learned
whom it stands for. If we do not have to learn its individual referent, we
presumably make use of a fundamental competence with which we are
already endowed, principally that of perceiving objects. The example is,
then, meant to refute the conventionalist view in favor of a perceptu-
alist view. On this perceptualist view, in order to know what is special
about pictures, we have to know what is special about the competence we
employ when treating something as a picture.
Let’s take a closer look at the example. We may assume in general that
we can identify an individual entity as individual only if we already know
something about it. If, for example, we want to apply a proper name cor-
rectly, we must either have already become acquainted with the corre-
sponding person or have obtained some knowledge of him or her from
169 Resemblance Reconceived

other sources—for instance, via description.6 However, this does not apply
to a passport photograph. Unlike the proper name, the picture gives us the
clues necessary to determine the person that the photograph refers to. The
clues come from the picture and only from the picture. We do not need
any additional information, nor do we need to know the person.7 In this
regard, the mechanisms by which a photograph and a proper name func-
tion are basically different, although they may be used similarly in certain
situations. (The case of twins is a special case I will discuss later on.)
If we accept the assumption that we get the information necessary to
correctly interpret a passport photograph exclusively from the physical
basis of the picture—the “sign vehicle” so to speak—we must also assume
that there is a nonconventional relationship between the picture and the
object depicted. This relation has the following form: in order to be the
referent of a picture, an object must have some relevant properties in
common with the sign vehicle.8 Naturally, this relation can be thought of
as a resemblance relation. No other relation so clearly suggests itself. The
resemblance theorist thus wishes to use this example to elicit an intuition:
that resemblance is the only plausible relationship allowing us to describe
the identification. On this basis, the resemblance relationship could be
regarded as a necessary property of figurative pictures. Of course, the
resemblance theorist does not need to defend such a strong version. The
theorist might take the resemblance criterion merely as a sufficient condi-
tion for distinguishing pictorial signs from linguistic ones. This weaker
version of the theory is the one I shall argue for in what follows. The inter-
pretation of pictures does not necessarily make use of this resemblance cri-
terion. There might be some pictures (or some aspects of pictures) that we
interpret differently. But whenever we employ the resemblance criterion,
the sign thus interpreted should be classified as a picture. This allows the
resemblance theorist to speak of a core area of depiction. My explication
is meant to be applicable at least to that area.

8.3 PICTURE THEORY AND RESEMBLANCE

My explication of the term “picture” consists mainly in introducing the


resemblance criterion in a rational way. “To introduce a criterion in a
rational way” means to argue step by step what this criterion is supposed
to do and why it should be adopted. This will at the very least provide us
with a better understanding of why and at what point the resemblance
criterion fails to do its job.
170 Klaus Sachs-Hombach

Before starting my explication, I would like to make a few method-


ological remarks. Although I assume that one can arrive at a full under-
standing of how pictures function only if one combines the findings of
different sciences, particularly of psychology and philosophy, these aspects
are on different methodical levels and should be kept apart. The philo-
sophical explication of the term “picture” formulates criteria that allow us
to classify something as a picture. This should be distinguished from the
question under what conditions an individual can apply such criteria. I
consider that question only briefly in section 8.4.
As I have already noted, I will proceed from a narrow concept of
pictures. As pictures in this sense, I will address only such entities that
are two-dimensional (and thus visually perceptible), artificial, and at least
semipermanent. The first criterion excludes mental and three-dimensional
pictures. The second criterion excludes so-called natural pictures, for
example reflected images. The last criterion excludes transient phenomena
such as clouds, in which we might see a landscape.
Pictures in this narrow sense are therefore paintings, drawings, or pho-
tographs. The broader phenomenon, which would include most of what I
have just excluded, might be termed “iconicity.” Given this specificity (to
which I should add the further restriction “figurative pictures”), I do not
intend to claim that we should use the term “picture” only in this narrow
sense. Instead, I restrict my explication to a core area in order to have a
solid starting point. In respect to this area, I shall introduce the resem-
blance criterion as a sufficient condition that should allow us to classify
certain signs as pictures. Here is a first draft version of the explication I
am going to set out and amend step by step:
(S1) An object is a picture if it is used as a sign and if it resembles what
it refers to.
This explication is of course in need of further explanation. The term
“sign” and the term “resemblance” are, as yet, undefined. Concerning the
term “sign” I consider a rather short explanation sufficient:
(S2) An object is a sign if we take it to stand for something else, that is,
if we assign a content to it, which, in many cases, allows us to determine
a referent.
Statement (S2) is not supposed to answer the question of what a picture
is. It only offers some distinctions that I take to be unproblematic within
semiotics. According to the semiotic triangle, (S2) proposes to distinguish
between the sign vehicle, the sign content, and the sign referent. In con-
171 Resemblance Reconceived

nection with (S1), the statement suggests that such a distinction also
makes sense for pictures. So pictures should be considered as normally
having a vehicle, a content, and a referent. The vehicle is a physical object
with properties, such as the visual ones of shape and color. The referent of
a picture is what the picture is taken to refer to: this can include all kinds
of objects and events—for example, the person we recognized on the road
after seeing the passport photograph. Finally, the content consists in those
properties that we consider to be relevant for representation and that we
relate to an object or an event. Thus the content normally provides us with
a procedure by which to determine the referent of the picture. The content
can also be characterized as an intentional object, which—so to speak—
we can see in the surface of the picture vehicle.9 A somewhat different but,
I think, equivalent formulation would be to say that the content is what
we see the relevant properties of the picture vehicle as. For example, we
might see a certain line on a canvas as the silhouette of mountains. The
process that constitutes a content should be termed “interpretation.”
As far as the distinction of vehicle, content, and referent is presupposed
in the following, pictures are signs by stipulation. Because this is the
meaning of the term “picture,” I would find it difficult to understand what
was meant if someone denied that pictures were signs. But one might
also say that in the following I am going to deal with pictures only insofar
as they are signs. Thus the question I want to answer is What criterion
allows us to distinguish pictorial from other kinds of signs, particularly
linguistic signs? I would like to argue that the resemblance criterion will
perform this function. In order to argue this, I am now going to introduce
the notion of resemblance, in a form initially independent of the picture
notion.
(S3) Objects resemble each other if they have, in relevant respects, essen-
tial properties in common.
Putting it this way, the term “resemblance” remains vague, because there
is nothing said about the meaning of “relevant” and “essential.” But this
reflects the way we use this term. Normally, we take resemblance to be
a gradual relation that applies to different objects only with regard to a
certain respect. Which respect should count as relevant depends on our
intentions and on the context. Compared to tables, quite different chairs
resemble each other, but compared to a variety of other chairs they might
not, because now size, color, or height are relevant. Also, what property
should be taken as essential depends, of course, on what respect is rele-
vant. Furthermore, it depends on how our perception works. We take, for
172 Klaus Sachs-Hombach

example, certain shapes to be typical as shapes of dogs or cats. An essen-


tial property is, therefore, a property typical of visual appearance that
normally helps us to classify something when we have seen it. How this
mechanism really works is an interesting empirical question. It might
be subject to cultural and historical influences. However, in order to be
more precise about what is taken as essential, one would need to have
greater empirical knowledge about how we classify different appearances.
Because I am not pursuing a psychological account, I will not go into
this problem. Within the framework of my philosophical explication, it is
sufficient to point out that such a psychological variable should be taken
into account. Whatever its outcome, it strongly suggests internalization of
the resemblance criterion. I have formulated this suggestion in the follow-
ing sentence:
(S4) Objects have essential properties in common if the perception of
the corresponding properties under a certain perspective is identical.
In adopting this internalized resemblance criterion, I abstain from the
possibility of explicating resemblance as a mathematically explicable rela-
tion between two sets.10 Instead of this (but without intending to criticize
it), I would like to take an intuitive notion of resemblance as a starting
point. According to this intuition, things are considered similar when they
seem so to us. Regarding the resemblance criterion as internalized this way
does, of course, not answer the question about which are the essential
properties, but it does suggest that one should discuss it on the percep-
tual level. Thus (S4) emphasizes the perceptual conditions under which
we experience something as similar, namely, when the perception of the
corresponding essential properties is identical. I think that this move
allows us to account for the variety of different styles more effectively.
Furthermore, it suggests why making something look realistic is not the
same as depicting something correctly from the geometrical point of view.
For example, in relation to a house, the front door should be painted
larger than it is, if it is to appear correctly.
Note that it is not required that the entire perception be identical. This
may apply to the special case of perceiving twins or illusionistic paintings
(trompe l’oeil). But, normally, we have no difficulty in distinguishing the
two things or events we consider similar. According to my explication, the
reason for this is that we refer only to the essential properties when we
experience something as similar to something else, whereas the simultane-
ous perception of the other properties enables us to be aware of their dif-
ferences. It is of course difficult to separate the two components of our
173 Resemblance Reconceived

perception, but that does not falsify my explication. Having internalized


the resemblance criterion in this way, we can now apply it to the sign rela-
tion. I will do this in two steps.
(S5) A sign resembles some other object if the sign vehicle and the other
object have, in relevant respects, essential properties in common.
Statement (S5) is directly linked to (S3). It simply says that the resem-
blance criterion also applies to objects used as sign vehicles, indepen-
dently of the sign relation. Statement (S5) does not really follow from
(S3), because, strictly speaking, only the sign vehicle resembles some other
object, whereas the notion of a sign includes at least a content. But I take
(S5) to reflect our normal way of talking and I propose that we actually
refer to the sign vehicle when we say that a sign resembles some other
object. I find this argument unproblematic insofar as signs necessarily
have a sign vehicle and can be considered as objects in this respect. But
putting it this way does not, of course, tell us very much. More important
is the second step, which is intended to characterize pictorial signs.
(S6) A sign is a picture if the sign vehicle and some other object have, in
relevant respects, essential properties in common (i.e., if they resemble
each other) and if these properties are constitutive for the interpretation
of the sign.
In order to use resemblance as a criterion to distinguish pictorial from
other kinds of signs, (S6) adds a restriction to the resemblance criterion.
It excludes in its second part those properties that other objects may also
have but which do not contribute to the content of the sign, that is, which
are irrelevant to its interpretation. Regarding something as a picture
makes irrelevant how heavy it is or what its back looks like. Therefore, in
order to be taken as a pictorial sign, it is not enough that the sign vehicle
has essential properties in common with some other object. These prop-
erties must also be regarded as relevant for the representational function
of the sign.
One could leave out the second part if one understood the expression
“in relevant respects” in (S6) as saying “in respects relevant for represen-
tation.” But because I have already suggested in (S3) that objects resem-
ble each other in principle in relevant respects only, for example, in respect
of color, the notion of relevance is already introduced in a broad sense. I
therefore prefer (S6) the way it is. It is more explicit about the need to
assume two different levels within resemblance theory. First, one deter-
mines relevant respects while comparing two objects and takes them to
174 Klaus Sachs-Hombach

resemble each other. Second, one determines the respects that are also
relevant for representation, that is, constitutive for interpretation. Thus,
“to take an object as a picture” means that we, sometimes automatically,
delimit the respects relevant for resemblance to the respects that are also
relevant for representation.
Compared to (S5), (S6) points out that there might be signs that are
similar to other objects without being pictures of these objects, because
some similarities are not relevant for depiction. Ruling out such cases,
for example, self-referential linguistic expressions, reference to the process
of interpretation implies that our classification of different kinds of signs
depends on how we use the signs. This implication seems acceptable,
because, depending on the context, we actually classify some objects as
pictures or as texts. In both cases we have to take the object as standing
for something else, that is, we have to assume a content and, if possible, a
referent, but the respects we consider relevant—and which partially deter-
mine the content—are not the same for pictures and text. In the case of
linguistic expressions, we do not, for example, consider color relevant. In
determining the properties that are relevant for representing something,
we therefore also determine what kind of sign we take this to be. In some
cases, mainly perspectival pictures that conform to Albertian principles,
this happens immediately and involuntarily. According to my explication,
this is exactly what one would expect, because resemblance is a gradual
relation: the greater the number of essential properties a picture has in
common with an other object, the more easily we recognize this object in
the picture. Often this also implies that more respects are relevant for
representation. Thus, perceiving the picture becomes more and more like
perceiving the object depicted itself. This case might be described more
adequately by (S7), which simply combines (S4) and (S6):
(S7) A sign is a picture if the perception of the essential properties that
the sign vehicle has in relevant respects is identical to the perception one
would have of the corresponding properties of some other object under a
certain perspective and if this perception is constitutive for the interpre-
tation of the sign.
Determining which properties we take to be relevant for representing
something is also helpful in distinguishing kinds of pictures. In a diagram,
for example, it is not important to know how thick a line is, whereas
this is normally relevant in a painting. Goodman has named this aspect
“repleteness.” In my explication, it is linked to the resemblance view.
Because the resemblance criterion is vague, that is, being determined in
175 Resemblance Reconceived

regards to certain respects only, it allows us to accommodate quite differ-


ent phenomena, which I have summarized under the heading of “iconic-
ity.” In photography, we regard most respects as relevant that are also
relevant in perceiving objects, whereas in diagrams we leave aside many
respects. “To leave respects aside” means that the picture is taken to
resemble some object only relative to the remaining respects. A diagram
does not tell us anything about what an object looks like.
But the more we abstract entire respects, the more we need to know
about the pictorial subsystem in order to be able to understand a picture
adequately. This leaves room for cultural influences. Here I am in agree-
ment with Ernst Gombrich’s position. According to Gombrich, the differ-
ent pictorial schemata in art history have been developed in respect of
particular functions the pictures were supposed to perform. Such a devel-
opment is only possible because the picture vehicle does not in itself deter-
mine which properties are relevant for the depiction, so it remains possible
to establish different respects as dominant. The pictures still resemble their
objects but only in regards to particular respects.

8.4 CONTENT AND REFERENCE

If my explication is consistent, it also indicates how we should understand


the relation between pictorial content and pictorial reference mentioned in
the example of the passport photograph. I now add some brief remarks
regarding this problem. According to the internalized resemblance crite-
rion, pictorial content is related to visual experience. It is therefore an
intentional phenomenon and certainly not identical with the pictorial ref-
erent. It consists in the perceived properties we consider essential to the
object depicted and which provide us with a criterion for determining the
referent. According to this criterion, the referent must be such that if per-
ceived, it would cause a perception of its essential properties identical to
the perception of the corresponding properties of the picture. This applies,
too, if you consider Goodman’s example of a photograph of a black horse
showing only a pale speck in the distance. Could this possibly be a picture
of a horse? Normally, of course, we would not take it to be a horse pic-
ture, but because we perceive horses in the far distance only as dots, this
example does not contradict the assumption that the pictorial content par-
tially determines the referent.
Pictorial content therefore provides us with a necessary condition for
determining the reference of a picture. But actually there are several imped-
iments to the determination of a referent, and this shows that pictorial
176 Klaus Sachs-Hombach

content cannot supply us with a sufficient condition. Pictorial content


always comprises a whole set of elements, without indicating which ele-
ment is referential. This effect is not confined to cases in which the picto-
rial content is ambiguous, where, for example, we see in certain lines either
a rabbit or a duck. In principle, we never know just from looking at the
picture vehicle which object is being referred to. This is no less true of a
passport photograph, because the subject of the photograph might have
an identical twin. In addition to pictorial content, the underlying causal
chain is here required for us to finally determine which twin is meant.
Reference is never, therefore, decided at the level of pictorial content
alone; but the content always gives us a set of elements. If we want to
know which particular element is referential, we normally need also to
know the context in which the picture is used.11
Another interesting case is the problem of different photographic prints.
We do not consider successive prints as pictures of each other. But would
this not follow from (S1), as the prints are signs and very similar to the
other prints? The answer is as follows: in principle, a print can be a picture
of another print, but normally we do not use prints in this way, because
no visual properties are emphasized that we consider essential for prints.
Consequently, we have no problem in accepting photorealist paintings as
pictures of photographs, because here precisely properties are emphasized
that are essential for photographs. If such hints are missing, we just take
those respects to be relevant that are also relevant in ordinary perception.
Now the basic notions related to the concept of a picture have all
been introduced. According to these notions, pictures are perception-
based signs. This means not only that we must perceive pictures as signs
(all signs must be perceived as such) but that their interpretation is in a
systematic manner bound to our perceptual competences. This allows us
to specify when someone is able to use an object as a picture. It presup-
poses a minimal understanding of how signs function and a rather com-
plex perceptual competence. “To take an object to be a picture,” then,
means that we perceive it as though we simultaneously perceived the
object itself in relevant respects. Normally, we remain aware that we per-
ceive the picture and not the object depicted, because the perception is
identical to the corresponding perception of the object only in relevant
respects.
Does this explication tell us much? Is it helpful in understanding a
painting? I think one should not expect too much from conceptual analy-
sis. To understand a particular painting is surely not equivalent to recog-
177 Resemblance Reconceived

nizing the subject depicted. Using a picture is a communicative act that


includes very different levels of meaning. We might use a picture to visu-
alize something, but we can also use it to express our feelings. Often pic-
tures symbolize abstract concepts, such as justice. Resemblance theory
thus reveals little or nothing about the meaning and value of famous paint-
ings. Nevertheless, what my explication shows is that our understanding
of pictures is grounded in perceptual competences. This helps to account
for our customary attitude toward pictures, for example, that we use them
to illustrate what things look like. Pictures call upon perceptual com-
petences that we normally employ without questioning their reliability.
Because this intermediation is very basic, our use of pictures cannot be
understood adequately without a grounding of the resemblance-theory
type.

Notes
1. See Lopes (1996, p. 11).
2. See Goodman (1968).
3. This view is sometimes ascribed to Gombrich (see Gombrich 1960 and 1982),
but the ascription is only partially correct. See Lopes (1992). A refined version of
the resemblance view can be found in Peacocke (1987).
4. For critical comments, see Scholz (1991).
5. My proposal is oriented toward the view Gombrich holds. See also Sachs-
Hombach (1999).
6. There is, of course, much more to be said about proper names. In order to con-
trast them with our intuitive resemblance theory, however, it is sufficient to point
to their conventional nature.
7. The claim I am considering here is termed “natural generativity” by Flint
Schier (see 1986, p. 43). See also the comments on this feature by Hopkins (1998,
p. 31 f.).
8. Having some properties in common with the sign vehicle is not, however, a
sufficient condition for being the referent of a picture. Particularly in photography,
we also regard the underlying causal relation between object and picture as essen-
tial. I will discuss this problem later.
9. For a deeper discussion of “seeing-in,” see Wollheim (1980) and Hopkins
(1998, p. 15 f.).
10. For such an explication, see Rehkämper (1995).
11. Pictures should thus be seen as analogous to general terms, not to proper
names. See Sachs-Hombach and Rekämper (1998, p. 124).
This page intentionally left blank
Chapter 9
What You See Is What You Get: The Problems of Linear Perspective

Klaus Rehkämper

9.1 INTRODUCTION

The two leading questions of this paper are, Is there just one way to see
the world? and, closely connected to that, Is there a correct way to repre-
sent the world pictorially? The plain and simple answer to both ques-
tions is, Yes, there is! Perspectival pictures represent the world correctly,
because the underlying theory describes correctly the way human beings
see the world as well as the making of correct pictorial representations.
The theory of perspective connects picture making and human vision.
Several arguments against this view, which is sketched very briefly, have
been formulated over the years. The three most prominent will be dis-
cussed in following: the curvilinearity argument, the immobile eye argu-
ment, and the argument that a picture drawn according to the rules of
linear perspective would not be accepted as natural. A closer examination
of these arguments shows that all three are not convincing and that they
also fail to show that the use of linear perspective in pictorial representa-
tions is merely a convention.

9.2 THE THEORY OF LINEAR PERSPECTIVE AND ITS CRITICS

Let me start with a few remarks about the different answers different sci-
entists have given over the years to the aforementioned questions. Euclid
was the first to describe the process of vision geometrically; he defended
the rectilinear propagation of light (a fact that no one has doubted and it
holds as long as we are talking about things happening on earth; here we
live in a Euclidean space). And Euclid also introduced the cone of vision.
During the Renaissance the rules of linear perspective—a system of rep-
resenting spatial relations—were discovered and described by Filippo
180 Klaus Rehkämper

Brunelleschi, Alberti, and Leonardo da Vinci, to name just a few. A


picture drawn in linear perspective—so the argument goes—is correct
because human vision is correctly described by these rules. A representa-
tional painting is like a window onto the real world, and the picture is the
window pane. All questions concerning the way in which to represent
space seemed to be solved; linear perspective was the answer. But at the
end of the nineteenth century, some new ideas arose. Linear perspective
taught that every straight line has to be represented by a straight line (or
in some extreme cases by a single point). Because the human retina is not
a plane (as pictures usually are) but a part of a sphere, straight lines—the
new dogma said—ought to be represented as curved lines. So it seemed
that linear perspective was not the correct and natural way of represent-
ing things; it was just one and not even the best way. The question of
whether there is a natural way of representing space or not was again open
for discussion.
Before I go any further in describing the history and the content of lin-
ear perspective and its rivals, I will try to clear up the different alternatives,
with the help of a quotation from a paper by Maurice Pirenne: “We must
decide between two main propositions: either (A) there is a natural system
of perspective corresponding to the way we see; or (B) there is no such sys-
tem, a number of different artificial systems being perhaps all equally
valid, or invalid. Further, if (A) is true, then either (Aa) Renaissance per-
spective is the natural system of perspective, or (Ab) the natural system
differs from Renaissance perspective” (Pirenne 1952/1953, p. 170).
This is a very good description of the existing alternatives. If there is a
natural way of representing things, it is either linear or another kind of
perspective, perhaps one that is not yet known; or it is all a matter of con-
vention. Convention means that every pictorial symbol system has to be
learned, and such a system can be chosen almost at random. There is no
inherent connection between the pictorial symbols we use and the things
represented.
Nelson Goodman in his book Languages of Art (Goodman 1968/1976)
attempts to defend this latter alternative, (B). In trying to represent spa-
tial relations he states, “The behavior of light sanctions neither our usual
nor any other way of rendering space; and perspective provides no absolute
or independent standard of fidelity” (ibid., p. 19).
Before I probe deeper into Goodman’s arguments, I shall quote another
source, the psychologist James J. Gibson: “From what I know of the per-
ceptual process, it does not seem reasonable to assert the use of perspec-
181 What You See Is What You Get

tive in paintings is merely a convention, to be used or discarded by the


painter as he chooses. Nor is it possible that new laws of geometrical per-
spective will be discovered, to overthrow the old ones.... There are no dif-
ferences among people in the basic way of seeing—that is by means of
light, and by the way of the rectilinear propagation of light. When the
artist transcribes what he sees upon a two-dimensional surface, he uses
perspective geometry, of necessity” (Gibson 1960, p. 227).
These quotations should be sufficient to prove that both alternatives (A)
and (B) can be found in the discussion about linear perspective. But what
are the arguments against the thesis that linear perspective gives the cor-
rect way of representing the things in the world? There are mainly two
arguments: (1) the curvilinearity argument, and (2) the immobile eye argu-
ment. But Nelson Goodman adds a third one: the argument that a picture
drawn according to the rules of linear perspective would not be accepted
as natural.

9.3 THE CURVILINEARITY ARGUMENT

I have mentioned argument 1 very briefly at the beginning of this chap-


ter. The retina is not a plane but a part of a sphere, and therefore it is
vaulted. Hence, straight lines are not seen straight but curved. And con-
sequently an artist, who wishes to represent what he or she sees, has
to paint straight lines curved. This argument was at first brought up by
the German mathematician Guido Hauck at the end of the nineteenth
century, but it became well known through the art historian Erwin
Panofsky’s famous article “Perspective as ‘symbolic’ form,” published
in 1927. Panofsky denies that the rules of Renaissance perspective hold
either for vision or for naturally representing spatial relations in a pic-
ture: “Finally, perspectival construction ignores the crucial circumstance
that this retinal image—entirely apart from its subsequent psychological
‘interpretation,’ and apart from the fact that the eyes move—is a projec-
tion not on a flat but on a curved surface. Thus already on this lowest, still
prepsychological level of facts there is a fundamental discrepancy between
‘reality’ and its construction. This is also true, of course, for the entirely
analogous operation of the camera” (Panofsky 1927/1991, p. 31).
And he continues arguing that “correct” representations of straight
lines have to be curved: “[F]or while perspective projects straight lines as
straight lines, our eye perceives them (from the center of projection) as
convex curves. A normal checkerboard pattern appears at close range to
182 Klaus Rehkämper

swell out in the form of a shield; an objectively curved checkerboard, by


the same token, will straighten itself out. The orthogonals of a building,
which in normal perspectival construction appear straight, would, if they
were to correspond to the factual retinal image, have to be drawn as
curves” (ibid., p. 32 f.). So the curvilinerarist will drop alternative (Aa)
from Pirenne’s description, but he has not to decide whether (Ab) or (B)
is true.
I postpone the discussion of this argument for when I try to rescue the
theory of linear perspective. But let me just ask two questions: Do artists
really to reproduce their retinal images in order to produce a correct and
natural representation? Should artists really draw the world as it appears
to them?

9.4 GOODMAN’S ARGUMENTS

Arguments 2 and 3 (from the end of section 9.2) can be found in Good-
man. At first he gives a correct summary of the basic assumptions of lin-
ear perspective: a picture drawn in correct perspective will, under specified
conditions, deliver to the eye a bundle of light rays matching that which
would be delivered by the object itself. This matching is a purely objec-
tive matter, measurable by instruments. And such matching constitutes
fidelity of representation; for because light rays are all the eye can receive
from either picture or object, identity in pattern of light rays must consti-
tute identity of appearence (Goodman 1968/1976, p. 11). Gibson would
agree with this summary; in his 1960 paper he states: “It is the light to the
eye that counts. . . . [O]ther things being equal, two identical instances of
stimulation must arouse the same percept” (Gibson 1960, pp. 219, 222).
Goodman takes his descripton as a basis for his arguments against the
theory of linear perspective. In the first of the two arguments, he tries to
show that the “specified conditions” required are too artificial and impos-
sible to achieve, an argument that already can be found in Panofsky’s arti-
cle. A picture drawn in linear perspective has to be looked at with a single
unmoving eye from a certain well defined-distance. Usually this is done by
looking through a peephole. The represented object itself must be viewed
through a peephole but not necessarily from the same distance and under
the same angle. First of all, according to Goodman, if the eye is not
allowed to move it becomes blind. Even if one allows the eye to move, “the
specified conditions of observation are grossly abnormal” (Goodman
1968/1976, p. 13). Under circumstances that are no more artificial than an
immobile eye—e.g., using suitably contrived lenses—one can wring out of
183 What You See Is What You Get

nearly every picture (i.e., pictures not drawn according to the rules of per-
spective) a pattern of light rays that matches. And so, he continues, even
if the patterns of stimulation are identical, “the same stimulus gives rise to
different visual experiences under different circumstances” (ibid., p. 14)—
A red light might mean “Stop” on the highway and “Port” at the sea.
Furthermore, we usually do not view pictures motionless through a peep-
hole but in a gallery, where we are free to walk around. So, the artist’s task
in representing an object is to decide what light rays, under gallery condi-
tions, will succeed in rendering what he sees. This is not a matter of copy-
ing but of conveying (ibid., p. 14). And Goodman closes this argument by
saying that “pictures in perspective, like any other pictures, have to be
read; and the ability to read has to be aquired” (ibid., p. 14).
For the sake of the first argument, Goodman has accepted the rules of
perspective, but in his second argument—argument 3 in section 9.2—he
attacks these rules. The main point now is handling the represention of
parallel lines. Goodman is not a curvilinearist, but he claims that pictures
that are accepted as natural do not obey the rules of perspective; in fact,
they disregard them. One of these rules, which he quotes, states that every
two parallel lines, which should be represented in a picture, have to be
drawn converging. This seems to conform with a statement of Panofsky’s,
about the rules of linear perspective: “[A]ll parallels, in whatever direction
they lie, have a common vanishing point” (Panofsky 1927/1991, p. 28).
But as a matter of fact, not all those lines are drawn in this correct way;
some stay parallel in perspectival pictures. And this is so only because of
convention, so the argument goes. Telephone poles bordering a street and
vanishing in the distance usually do not converge when represented in a
photograph but the street does; railroad tracks are another example of
this effect, as they are also drawn converging. But the edges of a facade
running upward from the eye are usually not represented as converging.
We use cameras with tilting backs and elevating lens-boards to avoid such
converging; we try to make vertical parallels come out parallel in a pic-
ture. And so Goodman concludes: “The rules of pictorial perspective no
more follow from the laws of optics than would rules calling for drawing
the tracks parallel and the poles converging. In diametric contradiction to
what Gibson says, the artist who wants to produce a spatial representa-
tion that the present day western eye will accept as faithful must defy the
‘laws of geometry’” (Goodman 1968/1976, p. 16).
It is clear from this short description of Goodman’s standpoint that he
argues for alternative (B). There is no natural system of representing
space: all such systems are based on convention and they have to be
184 Klaus Rehkämper

learned. Cubist or oriental paintings seem to be strange and unnatural to


us not because they are not drawn according to the rules of perspective
but because we are not used to them—we have not learned to read them
properly.

9.5 SOME REMARKS FOR RESCUE OF LINEAR PERSPECTIVE

Are these arguments convincing? In the following text, I will make some
remarks for rescue of linear perspective. The position I try to defend is
alternative (Aa): there is a natural system of representing spatial relations,
and this system is that of linear perspective. But this does not mean that
everything said against linear perspective is unreasonable, even if it is not
entirely right, if right means that linear perspective is shown to be wrong
by these arguments.
Let me give a short introduction to the theory of perspective. With this,
I will differentiate between two relations: (1) the relation of a picture
drawn according to the rules of perspective to the object depicted, and (2)
the relation that holds between such a picture and the beholder of that
picture. Relation 1 is a matter of mathematics, or to be more precise a
matter of projective geometry; relation 2 is where interpretation and there-
fore psychology come into play. Unfortunately these two very different
relations are often mixed up in the discussion about linear perspective.
What you need to produce a two-dimensional representation of a three-
dimensional scene in perspective is an object or an arrangment of objects,
G, a center of projection (often called the eyepoint), O, a cone or pyramid
of vision having its apex in O and its base at G, and a surface, B, inter-
secting the cone of vision. Now imagine (light) rays connecting the ob-
ject and the eyepoint. These rays are also intersected by the surface, and
the intersection points make up the two-dimensional representation of
that three-dimensional object relative to the eyepoint. It is important to
note that the surface intersecting the cone of vision need not be flat. It
might also be curved as in GB (figure 9.1). And you can also extend the
cone of vision beyond the center of projection into another cone (figure
9.2).
Intersecting the cone of vision not between G and O but behind O you
get a representation in perspective that is turned upside down and left-
right reversed (B2 or GB2). That is the way a camera obscura works.
Having a curved surface, for example, part of a sphere instead of a flat
plane, you get a picture that is similar to a retinal image. But as you can
185 What You See Is What You Get

Figure 9.1
The cone of vision with a flat and curved picture plane.

Figure 9.2
The cone of vision extended beyond the center of projection.
186 Klaus Rehkämper

see in figure 9.2, instead of having an object G to produce the pattern on


B2 or GB2, you can also use a picture in perspective of G such as B1 or
GB1. The results will be identical.
It all seems to be very simple. But I will give you an example of the
unholy mixture, not to differentiate between relations 1 and 2 I just men-
tioned. Although Panofsky in his aforementioned article correctly draws
the distinction between the “psychologically conditioned ‘visual image’”
and the “mechanically conditioned ‘retinal image’” (Panofsky 1927/1991,
p. 31), the latter of which is a correct picture in perspective, he is com-
pletely wrong in assuming “that the planar cross section of the visual
pyramid can pass for an adequate reproduction of our visual image”
(ibid., p. 29).1 And this assumption of Panofsky’s that a picture in per-
spective is an adequate reproduction of our visual image is shared by no
theorist of perspective.
Let me now return to the three arguments against the view that linear
perspective gives a correct and natural representation of spatial layouts. I
will start with the argument of curvilinearity. It is true, that the represen-
tation of a straight line on the retina is a curve, but this does not matter.
We do not see our retinal images. What we see is the real world, and pic-
tures are a part of this world. The task of the artist, while producing a rep-
resentation of the world, is not to paint her retinal image but to produce
a pattern of light on a two-dimensional surface that is identical to the pat-
tern emitted by the object viewed from a specific point. It is not the artist’s
subjective visual appearance that has to be represented. The theory of
linear perspective describes the relation of a picture to the world and not
the effect the picture has on an observer. If you draw a straight line curved
on the picture plane, you do not produce the same pattern of light. This
can easily be shown: a curved line on a window pane does not cover
completely a straight line in the world, so the patterns are not indistin-
guishable. There are of course occasions when a straight line is repre-
sented by a curved one, but only if the picture plane is also a curved
surface, for example, the ceiling of the sistine chapel or the movie screen
in a “Cinerama.” And these occasions agree completely with the theory of
perspective; the picture plane need not be flat.
The second argument—the argument from the immobile eye—is not in
itself convincing. It is true that the theory of perspective holds just for one
eye and the head has to be kept immobile, but the eye is allowed to move.
Picture and object are both entities of the world, and the eye can observe
the object in the same way it can observe the picture.
187 What You See Is What You Get

Figure 9.3
Different picture planes. O, eyepoint; B, picture plane; S, cone of vision; G, object.

The next point of discussion involves the different distances and angles
under which both object and picture have to be seen. As explained earlier,
what you need is an eye point, an object, a cone of vision having its apex
at the eye point and its base at the object, and—last but not least—a plane
intersecting the cone of vision usually somewhere between object and eye.
As mentioned before this plane need not be flat and it also need not be ver-
tical. All that counts is that the cone of vision is intersected by the plane.
From this, I think, it is clear that the distance from the eye point to the
object is usually greater than the distance from the eye point to picture.
And the angle under which the object is seen may differ from the angle of
eye point to picture. Suppose you a looking at the top of a tower, then the
main axis of your cone of vision meets the facade of the tower at an angle
less than 45°, but the picture plane may intersect the cone of vision at an
angle less than 90°.
That does not contradict the theory of perspective. But—and this leads
me to argument 3—Goodman is not right in assuming that all parallels of
the world always have to be drawn in a picture as converging. The paral-
lel edges of the facade are drawn converging in figure 9.3a but parallel in
figure 9.3b. This holds only for those parallels that are not parallel to the
picture plane; the others remain parallel. Or to rescue Panofsky, they have
a vanishing point in infinity. And if we use cameras with tilting backs, we
obey the rules of perspective. The only thing we are doing is changing the
angle under which the picture plane intersects the cone of vision and make
it parallel to the facade. And using lenses also does not defy these rules; it
is all in the range of physical optics.
I fully agree with Goodman that a stimulus can give rise to different
visual experiences under different circumstances. But this does not affect
188 Klaus Rehkämper

Figure 9.4
Floor or facade? (After Paul Klee.)

the theory of perspective, which holds for relation 1 described at the begin-
ning of this section. His argument belongs to the second relation, the
interpretation of the picture by someone who looks at it.
And while I am at it, Goodman also quotes the famous painter Paul
Klee to support his argument. Showing a drawing of Klee’s (imitated in
figure 9.4), he states: “As Klee remarks, the drawing looks quite normal
if taken as representing a floor but awry as representing a facade, even
though in the two cases parallels in the object represented recede equally
from the eye” (Goodman 1968/1976, p. 16 Fn. 16). But if one has a closer
look at Klee’s own description of the problem, it becomes obvious that he
is not supporting Goodman’s view.
Why is figure 9.4 wrong as a picture of a vertical facade? “It is not log-
ically wrong, because the lower windows are closer to the eye than the
upper windows, which in terms of perspective means ‘larger.’ As a repre-
sentation of a floor this perspective drawing would be at once accepted.
This picture is not logically wrong, but psychologically wrong. Because,
in the interests of maintaining its balance the animal wishes to see as
vertical all that is in reality vertical” (Klee 1925/1981, p. 31).2 Here again
Goodman but not Klee mixes up the two different relations. Arguing
against an interpretation, which is based on a psychological attitude, does
not affect the picture-object relation. Therefore it cannot show that per-
spective is merely a convention.
The last point I would like to mention is the problem that pictures
are usually looked at with two eyes and that pictures hang on walls. But
this does also not effect the theory of linear perspective. The theory
of linear perspective describes the relation between a picture and the
scene depicted under specified circumstances. Under these conditions
the theory obeys the laws of optics. And to borrow a phrase from Ernst
Gombrich: “Having achieved this aim, [the theory of perspective] makes
189 What You See Is What You Get

its bows and retires” (Gombrich 1960/1993, p. 217). The observer’s prob-
lem of how to read a picture while it hangs on the wall is not a problem of
the theory here under discussion. The conditions specified by the theory
are not fullfilled in this scenario. And if we leave the specified conditions
behind, then other problems crop up. In these cases we have to learn
to read the pictures. But in the case of perspectival pictures this is not
like learning a language, the symbols of which are chosen by convention.
These pictures have as a core the natural system of linear perspective—
a system that also describes correctly the way the human visual system
works—and that is why representational pictures of this kind are much
easier to read under “normal” conditions than any language is.

Notes
1. In the English edition Wood translates Sehbild as “visual image” in the first
quotation I have cited and “optical image” in the second. In doing this, he tones
down the contradiction but introduces a new, third term, which is, I believe, not
intended by Panofsky.
2. The original formulation used by Klee is “Warum ist Fig. 44 als Bild einer
senkrechten Hauswand falsch? Es ist nicht logisch falsch, denn die unteren
Fensterebreiten sind dem Auge näher als die oberen Fensterbreiten, was perspek-
tivisch “größer” bedeutet. Als Darstellung eines Bodens würde diese Perspektive
glatt akzeptiert. Dieses Bild ist also nicht logisch falsch, sondern psychologisch
falsch. Weil das Animal im Interesse seines Gleichgewichtes sämtliche Senkrechten
der Wirklichkeit auch als Senkrechte sehen will” (Klee 1925/1981, p. 31).
This page intentionally left blank
Chapter 10
Pictures of Perspective: Theory or Therapy?

Patrick Maynard

Dear Theo:
In my last letter you will have found a little sketch of that perspective frame I
mentioned. I just came back from the blacksmith, who made iron points for
the sticks and iron corners for the frame. It consists of two long stakes; the frame
can be attached to them either way with strong wooden pegs. So on the shore or
in the meadows or in the fields, one can look through it like a window. The ver-
tical lines and the horizontal lines of the frame and the diagonal lines and the
intersection, or else the division into squares, certainly give a few fundamental
pointers which help one make a solid drawing and which indicate the main lines
and proportions—at least for those who have some instinct for perspective and
some understanding of why and how the perspective causes an apparent change of
direction in the lines and change of size in the planes and in the whole mass.
Without this, the instrument is of little or no use, and looking through it makes
one dizzy. I think you can imagine how delightful it is to turn this “spy-hole”
frame on the sea, on the green meadows, or on the snowy fields in winter, or on
the fantastic network of thin and thick branches and trunks in autumn or on a
stormy sky. Long and continuous practice with it enables one to draw quick as
lightning—and, once the drawing is done firmly, to paint quick as lightning,
too. . . . The perspective frame is really a fine piece of workmanship. (van Gogh
1959, pp. 432–434)

Revealing shop talk about a technical device made for a stated practical
purpose, to be deployed by someone who knows how to use it: how dif-
ferent from what we are used to in the theoretical literature of perspective,
which as often as not can make one “dizzy.” For now we must put aside
the most interesting aspect of Vincent’s typically articulate and engaging
report: “pictorial space” in the sense of the composition of pictures, espe-
cially pictures of nature—including fantastic networks of branches and
the lie of the land and the clouds—so far removed from the cubes, crated
spheres and cylinders, checkerboard pavimenti of most theoretical presen-
tations. In this brief space we will stay with our project’s most simple,
192 Patrick Maynard

practical aims: that of identifying and addressing the bases of a cultur-


ally, generally confused state of ideas regarding pictorial perspective,
beginning with simple fallacies and concluding with more difficult matters.

10.1 REAL PERSPECTIVE: PERSPECTIVE USES

Luckily, our problems about linear perspective are not so bad, as they
exist more in the realm of theory than of practice. Let us begin with some
brief reminders about general practice, before considering the conceptual
problems.
As described by Vincent, his viewing frame was a device, a simple bit
of engineering. I suggest that what he used it for be thought of techno-
logically, also; indeed, that in our considerations of “pictorial space” we
recognize linear perspective itself as a kind of technology. What is a tech-
nology but an amplifier of our powers to do certain things?1 Looking at it
that way encourages us to be specific about perspective, to ask, do what
things? For his part, Vincent was clear: to make convincingly laid-out
depictive pictures from nature, whether drawings or paintings. How does
linear perspective allow this? Like any technology, by exploiting preex-
isting natural structures or forces. In this case the tapping into natural
powers is rather direct: linear perspective gathers overlap/occlusion and
foreshortening (which it shares with other projective systems) together
with its distinctive diminution with distance according to a simple rela-
tionship, all into one correlated system in which foreshortening is affected
by diminution and by other factors—that is, into a “space”—then feeds
the result into a visual system that is naturally tuned to using such infor-
mation. Zoom. Like most successful technologies, linear perspective devel-
oped from a variety of related dodges or practices, got consolidated and
then greatly elaborated at historical points, was diffused, and then—
again, like most successful technologies, such as the wheel, plow, water-
mill, printing press, automobile, computer—created its own markets and
rather changed the ends for which it was the means. This change of ends
is partly due to the fact that every technological amplifier is also a filter—
indeed a suppressor: get one form of power, lose another (which normally
we soon come to disvalue, even forget).
Perspective, like other technologies, has usually been mixed in with
its alleged rivals. Rather as we might call a Degas “a pastel,” despite the
other things it is in (and the underlying monotype), we usually call pic-
193 Pictures of Perspective

tures “perspective” when they are partly in perspective while also produced
through other shape- and space-rendering techniques. In this respect, per-
spective’s history is like that of a close pictorial relative, namely shade and
shadow techniques, which were also consolidated, systematized, theo-
rized, closely controlled, passed from specialized refinement to mixed use.
My next observation is that the listed components of linear perspective—
including foreshortening, overlap, diminution as a function of distance,
unified space—are frequently disassembled and stuck in with other sys-
tems. In general, it is good to approach linear perspective that way, in
terms of its components and their relations, rather than as monolith. As
we look around, we find what one might call “perspective patois,” even
“pidgin perspective,” in much of its familiar use in mass printing of images
today. Indeed I think that you will often find mainly oblique techniques
predominating there (with some orthogonal), though often inflected with
a bit of diminution—sometimes convergence of parallels—for effect: and
effect is, after all, the whole point of “pictures on a page.” What about the
photography which pervades these media? Is not linear perspective, as
we often hear, simply built into camera optics? Again, when we actually
look, I think we find similar mixes in commercial photography, notably
in advertising: the use of long lenses and closing up of backgrounds (often
stressing L-, fork, and arrow junctions), as the parallel systems, oblique
and orthogonal, predominate, with emphasis on object shape and texture
(so well rendered by photographs), with a residue of diminution. The
sense of a space, in terms of coordinated common vanishing points, is
rather rare in such work, perhaps because that would draw attention from
the surfaces that are being marketed to just space, which is not. Of course
one does see there, as in cinema and TV work, much strong perspective,
though again shapes on a page—including print and other graphics—
rather than “pictorial space” are the goal.

10.2 PERSPECTIVE SOPHISTRIES

Thus, very briefly, use. Now, what of conception, theory? I said that “we
call pictures perspective,” but who are we? Most people, including those
who ought to know better, seem rather vague about what perspective
and its main features actually are, typically conflating perspective with
a number of other projective techniques. For example, MoMA’s direc-
tor of painting describes an isometric cityscape as done by “perspectival
194 Patrick Maynard

plotting” and oblique sun-cast shadows rendered in an orthogonal pro-


jection as showing “perspectival deformations” (Varnedoe 1990, pp. 234,
238). A harmlessly loose usage, perhaps, unless it exemplifies pervasive
vagueness about that variety of projective drawing systems that guide the
most important drawings of the modern world: indeed those design draw-
ings that made a modern world possible. But the effects of this carelessness
get worse, fast. Consider that book on perspective said to be “the most
critically-acclaimed work of art history to have been published in France
through the second half of the twentieth century” (Carani 1998, p. 309),
elsewhere admired as of “vast learning,” and find, among its more than
400 pages, evidence that its author does not distinguish foreshortening
from diminution, does not remember that a picture can have other van-
ishing points besides the orthogonal, indeed cannot tell if a Renaissance
demonstration picture features an orthogonal vanishing point or not.2
The rave notices, from architecture as well as art faculty, are representa-
tive of those fields of higher learning. So what should we expect when we
turn to the more literary commentators?
Attitude follows ignorance. No one can be well versed in everything;
all of us go—one hopes corrigibly—on general impressions, dependent on
our own experiences to a point but also very much on general, nontechni-
cal treatments of the subject. This is where widely read, popular accounts
have influence, providing us glosses by which we think and talk about a
variety of matters. And on the subject of perspective of late, we all know
what that entails. Most people are probably unmotivated to know about
linear perspective. Perspective’s way into general discourse has therefore
not been via the sort of knowledge of those crucial drawing systems just
mentioned. Rather, linear perspective is usually introduced through “art,”
a conception that carries in our time a sense of great importance tinged
with unease. But regarding art, there was first perspective’s criticism as
a stultifying standard of the teachable part of an art education practice
now disdained; there was its alleged rejection by modern artists, beginning
with the Expressionists (though that situation is rather complex); there
were its alleged illusionistic tendencies, denying the materiality of the sup-
port. Most recently have come charges of Western perceptual imperialism:
“the hegemony of the eye,” “visual-ization,” and “the ideology of the vis-
ible linked to the Western tradition of systems centered on a single point.”3
As one observer has recently remarked, “postmodernism has grown out
of a neo-romantic tradition steeped in subjectivity and a general hostility
to the objectivity of perspective construction” (Topper 2000, p. 115).
195 Pictures of Perspective

Therefore these remarks belong to only the most recent version of what
Panofsky long ago called “the ultra-modern criticism of art,” in words still
worth citing:
It is now clear that the perspective view of space (not merely perspective con-
struction) could be attacked from two quite different sides: if Plato condemned it
in its modest beginnings because it distorts the “true sizes” of things and puts sub-
jective appearance and caprice in the place of reality and law, the ultra-modern
criticism of art makes exactly the opposite charge that it is the product of a nar-
row and narrowing rationalism. The ancient East, classical antiquity, the Middle
Ages, and every archaistic art, such as Botticelli’s, have—more or less completely—
rejected perspective because it seemed to bring into a world that goes beyond
or above the subjective something individualistic and casual; expressionism (for
recently there has been another swing of the pendulum) shunned it for just the
opposite reason: perspective confirms and preserves the remnant of objectivity
that even impressionism had still had to remove from the power of the individual
will toward form, that is, three-dimensional real space as such. But at bottom this
polarity is the two-fold aspect of one and the same thing and those objections aim
at one and the same point. (Panofsky 1924/1925, p. 18)

It may seem paradoxical that the common strategy of the more recent
“ultra-moderns” has been to attribute to perspective, among all drawing
systems, the most extreme form of subjectivity. Thus a very influential
popular book for nearly three decades states that “the convention of per-
spective, which is unique to Europe . . . centres everything on the eye
of the beholder.” “Every drawing or painting that used perspective,” it
continues, “proposed to the spectator that he was the unique centre of
the world” (Berger 1972, pp. 16, 18). Such opinions produced in cinema
theory ideas about what has been called “the ideology of the visible,”
with remarks such as the following: “[P]erspective-system images bind the
spectator in place, the suturing central position that is the sense of the
image, that sets its place (in place, the spectator completes that image as
its subject)” and “[T]he installation of the viewer as subject depends on
reserving for him or her, the reciprocal in front of the image of the van-
ishing point ‘behind’ it. . . . To launch an assault at the window [as de-
picted in Alfred Hitchcock’s The Birds] is, in turn, to assault the place of
the viewer [of it]; it is an act of aggression against the eye of the beholder
and the ‘I’ of the self-as-subject.”4

10.3 PERSPECTIVE FALLACIES, I: THE STANDARD HEURISTIC

Thus I remind you of muddles in much of what is taught and written


about perspective in our time. Practical remedy might include two steps.
196 Patrick Maynard

First, understanding of the rudiments of perspective and other projective


systems should be required of people getting graduate degrees in art his-
tory, cognitive psychology and related fields, and basic competence about
such matters should be required of literary people writing about the sub-
ject or people refereeing their manuscripts; also information about the
actual uses of these systems should be part of the curriculum. It is, after
all, not difficult to understand these systems in outline, and though it takes
a bit more time to learn about the facts of actual usage, most students
should find that “hands-on” study very interesting. Second, something
has to be done about a penchant for overgeneralization regarding per-
spective. In quoting those lines about viewers of Alfred Hitchcock’s film
The Birds, for example, I do not mean to reject the idea that a film—even
that cinema—might use perspective to make viewers experience their
viewing positions in such a way as to work some further effects on them.
Rather, I am criticizing the apriorism that attributes this effect to the mere
presence of perspective: that “perspective-system images bind the spec-
tator in place,” “suture” them to their seats, and so on.
What is the source of these excesses? My impression is that the under-
lying fallacies of perspective pervade the thought, writing, and teaching
of the actual curriculum I just endorsed: so much so that they have seri-
ously compromised our theoretical conceptions of pictorial depiction and
other kinds of graphic representation. The situation is so bad that we
may have somewhat to excuse the views I just quoted, as the responses of
those who, rightly perceiving danger, attempt in inarticulate rhetoric to
reply. Do I exaggerate? Let us begin by considering an admittedly extreme
position—and rather than picking old or easy targets, a sample from a
contemporary scientist of the human mind (from a source that, again, is
selling very well): “The picture exploits projection, the optical law that
makes perception such a hard problem. . . . Now, a picture is nothing but
a more convenient way of arranging matter so that it projects a pattern
identical to real objects” (Pinker 1997, pp. 215 f.). If we agree with that, it
is game over for “reconceiving pictorial space.” For if what a pictorial
depiction essentially is, is a perspective projection onto a surface (or the
inept, abbreviated or degenerate state of one), and also if what perspective
projection in depictions mainly is, is a way of representing space there,
there is no question of whether, when it comes to depicting space in a pic-
ture, perspective has first place.
However, pictures are not essentially in perspective or any other kind of
projections. Consider the converse: surface projections (perspective or
otherwise) as depictions. Although there exist in the universe perspective
197 Pictures of Perspective

and other kinds of projections—indeed uncountable real, physical projec-


tions: that is, wherever some kind of radiation gets partly blocked before
it hits a surface—a vanishingly small number of these are pictures or any
other kind of graphic representation. Turning that around again: there
has existed on this planet a much smaller but still uncountable number of
depictions and other kinds of graphic representations that are not even
partly perspectival, or even projective, not even in any inchoate or degen-
erate way. Although over the last demimillennium we have engineered
interesting intersection of these two huge classes (one of literally astro-
nomical proportions), there is not much more excuse for confusing that
intersection with the one class than for confusing it with the other. That
there is a persistent tendency to do so is something that calls for explana-
tion. What happened?
Following our technological approach, we can say that a very success-
ful technological means for (partly) producing images has been mistaken
for its end, depiction. Theoretically, having a well-systematized account of
perspective (along with other projective systems) but not even the begin-
nings of an account of depiction, cognitive inquirers, abhorring that vac-
uum, have stuffed the former in for the latter. This shows up generally, as
in the lines just quoted, or piecemeal, when findings about particular suc-
cessful perspectival effects in pictures are taken to show these to be consti-
tutive of depiction itself. Therefore I doubt that we will correctly conceive,
or reconceive, pictorial space until we correctly conceive depiction itself.
Short of offering a theory of depiction, we might at least keep in mind
some simple, familiar facts. We need to recall that even where perspective
is used to make depictions, it does not follow that the space defined by the
projection is the kind depicted in the picture. For example, when a movie
or TV camera produces, as it must, a perspective projection of part of
what is in front of it—that is, a series of props, including models, back
projections, and so on—the resulting depicted space is usually quite differ-
ent from the space relations that are projected: that is much the point of
using props and backgrounds. And we have all seen enough movie dream
sequences to know that the depicted space can even be nonperspectival.
Means is not the same as end: all this should seem as unsurprising as the
fact that it is possible to draw or to photograph plywood, cloth, and sty-
rofoam in order to depict stone, flesh, and metal.5 Why should anyone
have thought differently about the devices of linear perspective?
Therapeutically, we can trace the fallacy through a series of equivoca-
tions, which are conveniently gathered, it so happens, in the book that
coined the expression “linear perspective”: Brook Taylor’s Perspective,
198 Patrick Maynard

New Principles of Linear Perspective, whose first definition repeats its


subtitle as follows: “Linear perspective is the Art of describing exactly,
on a Plane Surface, the Representations of any given Objects,” which gets
parsed immediately as follows: “let the Reader consider, that a Picture
drawn in the utmost degree of Perfection, and placed in proper Position,
ought so to appear to the Spectator, that he should not be able to distin-
guish what is there represented, from the real original Objects actually
placed where they are represented to be” (Taylor 1719, p. 2; see also
Anderson 1992). Stop there and notice the equivocation on the very gen-
eral term “representation.” Taylor began by talking about a way of pro-
jecting 3-D patterns onto 2-D, called such projection “representation,”
then substituted the word “picture” (because pictures are often called
“representations”): voilà, a projection is a depictive picture. This equivo-
cal slide is now the slicker for being so well worn, as projections in cam-
eras and in eyes have long been loosely termed “pictures.”6 But, as just
stated, no mere projection onto a surface, whether actual or—as in the
case of nearly all expositions—fantastic and merely imagined, is either
necessary (not Taylor’s assumption) or sufficient (Taylor’s assumption)
for pictorial depiction of anything: a fortiori not for objects actually or
just notionally projected.
Compare a closely related possible equivocation: cast shadows are
very like pictures; also pictures can be made very effectively by use of cast
shadows—there are important technologies for this, such as X rays and
“magic shadows.” Therefore, why not call all pictures—or at least ones
made from shadows—“shadows,” or at least try to understand how we see
pictures by how we see shadows, or try to explain the pictures that are not
made by shadows as derivatives from the ones that are? This is clearly fal-
lacious: of course projections and shadows become pictorial only when
they are used in certain ways, in ways notably connected with visual per-
ception and imagining, as with hand shadows. The word “used” should
immediately put an end to a priori assumptions. Yet one of the most basic,
amazing, fallacies of most theorizing about perspective is the neglect of
difference between a tool (perspective) by which a picture might be made,
actually or theoretically, and the spatial characteristics that the picture
depicts or represents.
To be fair, Taylor and his descendants in viewpoint typically do indi-
cate one perceptual use, even include perceptual constraints. But that
narrow, artificial use is usually, as Taylor wrote, perceptual indistin-
guishability: you do a picture like that in order to produce a situation that
199 Pictures of Perspective

Figure 10.1
Faces in strange places: memento mori soup.

is visually indistinguishable from what another situation would be. Not a


promising description of depictive use, but one that has wide appeal—
even driving recent image technologies. As for the constraints, whereas the
old ones took account of what Taylor called “proper Position,” we find
contemporary accounts needing to append “post-Ames” remarks about
what “the brain is biased to recognize” (Pinker 1997, p. 215), or what
“excite neural circuits that would be excited by” entities that mattered
in the course of evolution (Shepard 1990, pp. 187, 198)—given that even
geometric projection is radically nondeterminative of depiction. (I am
grateful to Jan Koenderink for his ingenious subversion of the famous
Taylor diagram.) Such recent constraint qualifications turn out to be rather
more than qualifications. Indeed, anything such as a brain bias or “neural
circuit” approach to depiction dooms the projective paradigm for depic-
tion, by letting in facial, gestalt, movement, textural, and tactual factors—
a mass of other nonprojective inciters to visual imagining (figures 10.1,
10.2, and 10.3).
Therapy for theory would not be complete without noting another
pervasive fallacy expressed by Taylor’s account: a reciprocal one. Just as
200 Patrick Maynard

Figure 10.2
Faces in strange places: birch eyes.

theoretical conceptions of pictorial depiction have been confused by stan-


dard presentations of linear perspective, conceptions of perspective have
in turn been muddled by their pictorial applications. The theoretical meet-
ing of perspective and depiction has thus been a road accident of ideas in
which all parties’ heads have been banged. As a method of algebraic or
geometrical projection, perspective has nothing to say about vision. Also
it does not mention light. Nevertheless, in its standard heuristic presenta-
tions, when we are getting ready to think about its pictorial uses, it is usu-
ally explained in terms of highly theoretical, not real, light rays (typically
imagined to go clean through the wall or whatever surface the picture is
201 Pictures of Perspective

Figure 10.3
Natasha toothbrush.
202 Patrick Maynard

to be drawn on). Next, crucially: further anticipating the model—that is,


application to pictures—the projective-point variable for these rays is usu-
ally decorated by a picture of an eye, as in Taylor. Obviously, this eye has
nothing to do with the projection, considered geometrically or even light-
ray fantastically. Yet the problem of decorating the diagram with such
images is not simply one of what is called “chart junk”; the problem is
what has perhaps proved the worst source of confusion so far. And that is
the fallacy that (as we saw early on) linear perspective, as used in pictures,
automatically, inevitably, introduces subjectivity. That fallacy deserves
special treatment.
I have already suggested what misunderstanding of linear perspective
has done to the understanding of, and attitudes about, perspective among
those who are unclear about what linear perspective is. Following is a
recent report from someone who has read more of this material: “Mistrust
and indignation before the crass theatrics, the arrogance, the pedantry
and scientism of the so-called legitimately constructed perspectival picture
have become the norm among a wide range of theorists and historians.
Perspective is treated not merely as a severe distortion of the actual fact of
seeing. . . . Today perspective is treated as a complete abstraction . . . [even
as something] irrevocably ‘bourgeois’ or ‘ideological’” (Wood 1995,
p. 680). I also offered evidence that although, as our source reports, there
are other grounds for this resentment, much of it is owing to the alleged
egocentricism of perspective, as presupposing, or imposing, a subjective
viewing eye, so that, as one of our influential sources puts it, “the visible
world is arranged for the spectator as the universe was once thought to be
arranged for God” (Berger 1972, p. 18).
Confident rhetoric—or a cry for help? Once again, perhaps we should
not (much) blame the sources for these reactions, when even the experts
on the subject standardly present perspective constructions in pictures as
indeed made for a single privileged viewpoint, and as intelligible only by
reference to that viewpoint. Consider, for example, yet another influential
expert’s overgeneralization: “[An occluded]-surface perspective view—
whether photographed, hand-constructed, or computed—implies an eye-
witness. It suggests that there was an observer at the station point, and it
puts us precisely in the shoes of that invisible observer. . . . We are con-
fronted with the paradox of disembodied viewing—a paradox that every
photograph and every consistently constructed perspective contains”
(Mitchell 1992, p. 134). “Every photograph”—like those on packages,
such as cereal boxes? To be sure, this is again an extreme expression—
203 Pictures of Perspective

though, after all, by a computer graphics specialist. Still, what many per-
spective theorists normally say is not far off that. No point to a parade of
quotations here: we know that the standard heuristic exposition of lin-
ear perspective is simply, without argument, to identify the perspective
construction point as the picture’s viewer’s point and to take that quite
literally. This is sometimes even expressed (again by one of our quoted
experts) as “the viewing position of the artist” or “the artist’s unique van-
tage point” (Shepard 1990, pp. 191, 196 f.), sometimes as “where the artist
stood”—as though perhaps Masaccio had stood before the Trinity, or
Raphael behind Saint Paul preaching in Athens—or at least behind an
actor on a set, and so on! Part of the pity of this amazingly careless way
of speaking, thinking, is that—just as viewer position in cinema is some-
times part of the content of the experience—in some cases it does matter
that the artist was at that point, just as it sometimes does matter for the
content of a picture that we understand there to have been a viewer at a
more-or-less precise point, just as it sometimes does matter that we view
the picture from near the construction point. Sometimes these concerns
matter, but, in the overgeneralized, a priori fog round the topic, such
“finer” points of picture meaning are soon lost.7

10.4 PERSPECTIVE FALLACIES, II: VIEWS

Although I have tried to indicate the wide influence of the positions


reviewed here so far, it might be objected that, from Taylor on, they have
been extreme or carelessly stated. And, given that linear perspective is,
surely, through its exploitation of natural resources, a highly successful
technology for spatial rendering—thus rejecting the accounts of some
extreme current relativisms—that must have something to do with linear
perspective’s pictorial matchup to the light geometry of environmental
vision. This being true, the question is what to make of it, and the first
thing to do is to put our topic itself into useful “perspective.” So, as our
next reminder, consider the most loose, general meaning of the word
“perspective” as “point of view.” Dictionaries tell us that “perspective”
means prospètto, so, not “seeing through,” as Albrecht Dürer thought, but
“prospect,” or “view.”8 This suggests seeing from a distance, which has
two general connotations. First, we have perspective on matters when we
are not too close to them but, by standing back, can better understand
them in context. Second, we realize how things may appear different
from different perspectives or standpoints. In short, unless we are radical
204 Patrick Maynard

Figure 10.4

relativists, we will hold that some perspectives (in the second sense) afford
better perspectives (in the first sense) than do others—although, if we are
reasonable, we will also be willing to consider things from others’ faulty
perspectives. But what has this to do with pictorial perspective, except as
the agency that brought these usages into modern languages? One appeal
of a perspective paradigm for all depictions may have been that linear per-
spective renders what are considered views, and visually effective pictures
of any kind usually show views of things.9 That this is a constraint on
effective pictures is hardly surprising, because we often recognize things by
perspectival contours. Indeed, the famous word eidos, developed by Greek
philosophers and reworked by Latin as “form,” meant—and this root
meaning persisted through the theoretical elaborations—that about the
look, notably of the shape, of a thing that allows us visually to recognize
it. Consider, for example, a clear photograph that does not present such a
view (figure 10.4). When pictorial perspective does provide effective views,
there is nothing peculiar to it, and a “peephole” need no more be involved
in linear perspective than in any other method.10 Not only are most depic-
tive drawing systems designed to exploit such views, perspective has weak-
nesses as well as strengths in that regard. Such are the systems’ trade-offs,
or “trades”—one motive for that practical use of mixed devices discussed
at the outset.
205 Pictures of Perspective

To consider this point more closely, we need to consider views more


closely. So far we have been considering views of things (such as spaghetti
noodles in a plastic container), but people observe views in a literally
broader sense, which includes groups of objects. They get a room with a
view, admire the view. They find the view’s been spoiled by developers or
a volcano, or is not so good as if you’d just climb a few more steps; they
turn off the highway at posted scenic viewpoints. Indeed, there was once
a market for paintings (for example by Canaletto), called vedute—usually
in perspective—of views; and today tourism, whole economies, depend
on views. Such views are of course relative—perspectively relative, even
relative to humans—but they are not subjective, nor do they imply any
viewer. “The view from there is terrific” is on a par with “the water’s fine”
(that is, for normal swimmers), which evokes no ghostly bather or privi-
leged epidermis. With views of scenes, as with views of objects, it needs to
be emphasized that, contrary to popular claims, there are long traditions,
in literature as well as visual art—such as that of China—of vividly cap-
turing precisely such views, as views, innocent of any perspective concep-
tion of the job.
Recalling our technological context, when systematized linear perspec-
tive did come along, it joined an old family of view-depictions—to be sure,
with considerable jostling—bringing specially big advantages regard-
ing views, in both the object and the broader sense of “view.” But it also
brought disadvantages; such are the trades: every technological amplifier
being also a filter. Although perspective somewhat changed the pictorial
end, as people learned how to work it, and to see with it, it has joined the
general bag of depictive tricks. Most people are rather vague, looking at
perspective pictures—including their own photographs—about the loca-
tions for the views, even when they look at such pictures as views, which
is far from normally the case. Just as well: in most real cases (particularly
away from carpentered subject matter), it is impossible to retrieve the con-
struction points. In this regard people tend to treat perspective pictures
so like any others that, once again, the cases in which they do not so treat
them have special meanings.
Thus perspective “view,” as what is shown in pictures. But, oddly, much
of the controversy about picture perspective has to do with another topic:
our views of pictures. In other words, there is much recent interest in per-
spective views of perspective pictures themselves, that is, in how they work
when we see them (as we normally do) at a variety of slants and distances.
This interest is usually based on an assumption—call it “the coincidence
assumption”—that, in order to work, perspective pictures would need to
206 Patrick Maynard

be seen from their construction points. (Because such theorists usually


count parallel systems such as the engineers’ and architects’ orthogonal,
oblique, and axonometric as limiting cases of perspective, they should also
wonder why we would not prefer looking at such drawings at such remote
distances that we need binoculars.) The coincidence assumption, which
has a long history, is of course well refuted by normal practice, from
which we might draw the following analogy. Previous to the 1870s, when
articulate utterances were heard, one could assume that they were heard
from a position fairly close to the source, in time and in space; such were
our physical limitations. Technology changed that, for phonographs and
telephones freed us of those limits: thus “his master’s voice.” Previous to
the 1440s, when anyone had the experience of seeing a thing or view in
perspective, one could assume that this was from a position at the optical
construction point. Technology of course changed that, too, for linear
perspective technology (as used in pictures, as well as in stage settings,
architecture, gardens, etc.) freed us of that restriction, as well. That it
turns out that perspectives still work when viewed away from their con-
struction points does not mean that they work the same way or even as
well, but the same can be said about telephone, voice recording, and so on.
Nevertheless, they all do work, and it is, after all, typical of technologies
to loosen natural restrictions.
Frankie, looking at her snapshots of Johnny and herself, has no notion
that, given the lens on her 35 mm point-and-shoot, the projection point
for these 4˝_6˝s would be along a solid perpendicular to its middle, at
exactly the length of the ballpoint pen with which she just wrote on
the back—a viewing distance strictly for the nearsighted. But (our next
reminder), Frankie, like everybody else, tends to look at all pictures from
around the middle, roughly on a perpendicular—that is, when she can.
Although like everyone else, she usually settles for somewhat angled views,
close enough that she can see it all. Frankie also does this with the news-
paper (even when Johnny is reading it), with train timetables (despite
slow-reading backpackers planted in the middle), and with the titles and
printed credits on cinema and TV screens, when she is seated off to one
side. I recall a young woman intently viewing the Rijksmuseum “Golden
Age” show from a wheelchair, the pictures well above her head, in gallery
after gallery, twenty-one in all. Theorists may wonder how she could have
done this with all those perspective pictures, but the same issue arose if she
went to the Kurt Schwitters show down the street at the Stedelijk. Leave
it to those so gripped by the a priori projection-point heuristic that they
207 Pictures of Perspective

have invented a special research project for perspective. Some favor deft
trigonometric extrapolation of the projection point. But, as illustrated by
Frankie and the photo, not only will people not know what we are talking
about here, it is implausible to suppose that their visual systems have
much motive, or means, to do so. As we watch movies, flip magazine
pages, and encounter a fast flux of camera positions and focal lengths—in
magazines, of cropping formats—we show scant tendency to adjust our
viewing distances. The sense of close-up and long-shot certainly do regis-
ter, but that so-called close-up may have been taken some distance off
with a telephoto lens, to keep the background blurred, whereas the wide-
angle lens that was used close in mainly distorts features so as to make
faces look ugly or menacing.
The argument is not that it is vain for perceptual theory to investigate
generally how we reckon shapes on or of surfaces. After all, most of the
surfaces we work on or walk on meet vision at acute angles. Getting the
three blocks from the Rijksmuseum to the Stedelijk along Paulus
Potterstraat to see the Schwitters would involve a lot of estimating in that
regard—probably harder at a child’s or wheelchair’s lower perspective.
My worry is, again, about what is behind the assumption, without argu-
ment, that optical inclination poses a particular problem regarding per-
spective pictures. Perspective is capable of winning its way empirically
as a method of producing pictorial space (and other things) and has no
need for the standard stacked decks—or opening, a priori, sleights of
hand.

10.5 ANOTHER SLANT ON PICTURES

By now, real treatment of our cultural picture-perspective perplexity may


be on the way: not through words like these, but where we began, through
actual practice. Everyday experience with shifting small cameras and now
camcorders may be doing it. If the old geometric heuristic has been so
badly misleading, is it too much to hope that, as algebraic computer draw-
ing takes over, the old heuristic may lose its theoretical grip? Could it be
that alternative wireframe algorithms, computer translations, back and
forth, between x, y, and z coordinate-triples and x, y projection-pairs will
spread graphic literacy, as the division of the x and y coordinates by the
z, which characterizes perspective, may be understood as one effective
option, without having to imagine fictive light rays, an intermediate pro-
jection plane, and so on?
208 Patrick Maynard

I stress “back and forth” because I perceive another problem with that
heuristic beyond what I have indicated. It has continually—unwittingly,
to be sure, but influentially—imbued thinking with the prejudice that
drawing perspectively, or in any way projectively, has not only the con-
notation of passive, mechanical skills of projecting (“a machine could
do it”) but something deeper, which seems never articulated: the deriv-
ative connotation that underpins that heuristic, that the 3-D subject is
somehow given, whereas the drawing is gotten from it. Thus the stan-
dard talk about a “real world” or “real object” (mostly fictional) and “its”
picture-projection. If this connotation did not actively depress projective
drawing, it did little to help such drawing in its recent times of trouble.
Ironically, what actually happens is often the reverse; as remarked, most
of the real things that define the modern world (that is, all the artifacts in
it) are made from projective drawings, rather than the other way around.
I doubt that one will achieve a satisfactory account of the depiction of
space in pictures unless one has a good general account of depiction itself.
Not having attempted that, I close with two critical remarks in that direc-
tion, which are more radical than what I have argued so far. First, briefly,
insofar as the depictive challenge is spatial, why assume that this is neces-
sarily a matter of going between 2-D and 3-D, when 0-D, 1-D, 2-D, and
3-D are all involved, even in visual depiction? As visual depiction of space
is by no means confined to pictures, I suggest questioning the focus on
pictures that has dominated discussion—in isolation from reliefs, masks,
carving, sculpture, and so forth—even when talking narrowly about per-
spective. (Remember the meaning of three old words, arti del disegno.)11
Second, the idea “insofar as depiction is spatial” has been assumed, but
must it be? Far more radically, I suggest that we are unlikely to gain an
adequate conception of any kind of depiction so long as we think of it
even as an essentially spatial affair—less so, as long as we accept the stan-
dard idea that “the basic trick of pictures” (to quote one outstanding art
historian of our time) is that of “marking a flat plane to suggest the three-
dimensional” (Baxandall 1985, p. 106). Here is not the place to challenge
a dominant space-based conception of vision that reinforces the dominant
space-emphasizing approaches to visual depiction, which has less to say
about substance, process, movement.12 Still, as a “simple” example, the
earlier physiognomic examples should have already suggested what this
wonderful, eidos-capturing, 3,000-year-old, double-necked pot painting
shows: that not only are pictures not essentially flat or even plane, but that
we need to think whether the basic challenge of pictorial depiction has to
209 Pictures of Perspective

Figure 10.5
Stirrup jar: Octopi and fish. Mycenean, terracotta, c. 1200–1125 B.C., h. 10 41 inches.
The Metropolitan Museum of Modern Art, Louisa Eldridge McBurney Gift Fund,
1953. (53.11.6)

be dimensional—even spatial at all (figure 10.5). I point to a couple of


features of this example, regarding the theme of “pictorial space.”
Not only is the surface not flat; the painting uses the (double) curvature
of the pot to swell out the form of the cephalopod: indeed, its eyes big, the
startled animal seems to swell out at us. I mean swelling (a process), for
note how the vertical curvature changes, as the arms bend back in space,
slowly coiling in sequential rhythm. We also sense it settling again, for this
“still” picture is always in motion. And there is something else to notice:
how the image pays the pot back, for it decorates the pot it is painted on.
Its swelling enhances our experience of the convexity of the pot, as its
210 Patrick Maynard

motion out and up emphasizes our sense of the pot rising and expanding
through belly to shoulder, then closing through handles and lips, as the
fish spring up from the foot.
That such decorative depiction forms a central kind of depiction seems
rather lost on modern theorists, misled by a few centuries’ dominance of
technologies of “unlocated” depictions—that is, of images put on sur-
faces made for just that purpose and devoid of intrinsic interest—such as
we find in cinema and TV screens. Technologies of photo reproduction
convey the impression that such is the historical norm. And those who
take decorative depiction as marginal to issues of pictorial space might
recall that a favorite example of those drawn to the topic, Andrea Pozzo’s
painted ceiling for the nave of the Chiesa di Sant’ Ignazio in Rome
(1688–1694), is essentially—not coincidentally—a decorative depiction.
That ceiling was not just a convenient ground on which Pozzo could place
his images. As we began with technological uses of perspective, let us
recall that here linear perspective’s primary function is to promote our
imagining, of the very nave in which we stand, that its walls rise much
higher above us and are occupied by diverse figures in motion, that its roof
is open to the clouds, that the saint is visible above us, carried up by
angels. As decorative, the image is more than what Maurice Pirenne calls
Pozzo’s “imaginary structure” (Pirenne 1970, p. 79), for (and this hap-
pens, too, with some restaurant decor) its practical purpose is to have us
imagine things of that actual architecture—and thereby of ourselves, as
well. Through attention to such cases of perspective and other methods in
actual use we should find improved bases for an understanding of visual
depiction, which can guide our research into pictorial space.

Notes
1. I have taken this approach to photography, for example in Maynard (1997a,
1997b), and to perspective in Maynard (1994).
2. See Damisch (1994, pp. 63, 107, 158, etc.). I make further comment in review
of the book in Maynard (1997b, pp. 84–85). The quotations are from editorial
reviews available on booksellers’ Websites.
3. This citation is from Noël Carroll (1988, p. 128). For more careful argument of
the material in sections 10.2 and 10.3, see Maynard (1996, pp. 23–40).
4. These citations are also from Carroll (1988, pp. 128 f.).
5. It is even a fallacy to go from “x is a depictively essential constituent of (pic-
ture) P” to “P depicts something as x,” because that fails for such values of “x” as
being a painting, being monochrome, and so forth.
211 Pictures of Perspective

6. This does not necessarily commit the fallacy we began with, that of claiming
that all pictures are essentially projections. Notice how Henry Talbot could only
invent photography, a more recent source of the projection/picture mix-up, by
fighting off that very tendency: “The picture [sic]” from his camera, he wrote,
“divested of the ideas which accompany it, and considered only in its ultimate
nature, is but a succession or variety of stronger lights thrown upon one part of
the paper, and deeper shadows on another” (Talbot, 1844–1846/1969). If Talbot
meant his camera lucida, where the image is virtual, there would not have been
lights upon the paper.
7. These careless, common phrases might be revealing of something: maybe
drafters’ often assuming their pictures’ construction points when doing observa-
tional drawings, in order to “eyeball” the perspective, has encouraged the idea that
viewers need to do the same. An odd inference, in any case: it is like saying that
because roofers have to go on the roof to shingle it, you have to go up there, too,
to see the roof.
8. Erwin Panofsky noted Dürer’s error here in Panofsky (1924/1925): “[D]ies
‘lateinisch Wort,’ das schon bei Boethius vorkommt, ursprünglich einen so präg-
nanten Sinn gar nicht besessen zu haben scheint” (p. 258).
9. For an account of “effective pictures” and “views,” see Willats (1997, pp. 22 f.,
207, 219).
10. As is claimed in Shepard: “[E]ach picture of a three-dimensional scene carries
its own peephole” (1990, p. 122, repeated on p. 196). The photograph shows the
top of a spaghetti tube.
11. As Vasari’s phrase arti del disegno indicates, sharp distinctions among paint-
ing and sculpture, architecture, and relief are recent. To call pictures 2-D media is
somewhat misleading, as there, with 0-D, 1-D, and 2-D marks, we depict not just
3-D but also 0-, 1-, and 2-D situations. In 3-D sculpture and architecture we depict
the same, using 0-D junctions, 1-D contours, 2-D shapes and expanses, and so on.
It is also possible to depict in 1-D media.
12. In Maynard (2001), I hazard remarks about biological and technological vision
as aimed more at process and motion than at objects and spatial forms. Ernst
Gombrich is one leading theorist of depiction who has confessed an overemphasis
on space at the expense of representation of transparency and life (Gombrich
1987, p. 16).
This page intentionally left blank
PART III
THE NATURE AND STRUCTURE OF RECONCEIVED
PICTORIAL SPACE
This page intentionally left blank
Chapter 11
Reconceiving Perceptual Space

James E. Cutting

11.1 INTRODUCTION

How do we perceive the space in pictures? In answering this question


theorists typically consider standard photographs and other representa-
tional images, such as architectural drawings, engravings, and paintings
in linear perspective. Adding motion augments this domain, including
cinema and computer-graphic sequences. But cartoons, caricatures, and
much of twentieth-century art and photography are seldom overtly con-
sidered. This, I think, is a mistake; without a broader perspective on pic-
tures one is lulled into thinking too metrically, about pictures and about
the real world as well.
Certain philosophers and psychologists have spilled a lot of ink dis-
cussing pictures and the space depicted within them; other philosophers
and typically other psychologists have done the same about perceived
space in the world around us. What I propose is that our perception of
these spaces is done in pretty much the same way, that neither are guar-
anteed Euclidean but that they are built upon available information. In
essence my chapter is on the promise of a constructive and cooperative
measurement-theoretic approach to all perceived space. The more infor-
mation, the more constraints, the more any perceived space will incre-
mentally approach a Euclidean ideal.

A Précis of Measurement Theory


After his broadscale survey of how scientists measure their subject mat-
ters, S. S. Stevens (1951) reported that we measure the world in four
ways—using nominal, ordinal, interval, and ratio (or metric) scales (see
also Luce and Krumhansl 1988). Nominal scales name, or simply catego-
rize, differences. Such categorization is always the beginning of science,
216 James E. Cutting

and taxonomies (which are nested nominal scales) remains critically impor-
tant in biology today. Ordinal scales categorize and order, but they say
nothing about distances between what is ordered. The division of U.S.
college students into freshmen, sophomores, juniors, and seniors is an
example, for this only gives a rough idea of their “distance” from gradua-
tion requirements. Interval scales categorize, order, and provide true dis-
tances, but there is no true zero on such a scale. Thus, one can say that
the distance between 5°C and 10°C is the same as that between 10°C
and 15°C, but one cannot say that 10°C is twice as warm as 5°C. Finally,
ratio scales categorize, order, provide true distances, and have a true zero.
Thus, one can say that 1 m is half as long as 2 m. In doing psychophysics,
for example, one often manipulates a physical variable along a ratio
scale, and records some psychological variable that is almost surely only
ordinal.
Stevens’s (1951) classification system of the four scales is itself ordinal.
That is, one cannot really know the “distance” between nominal and ordi-
nal scales, for example, as compared to interval and ratio. However true
this is logically, one can nonetheless state psychologically that through
various considerations one may find that ordinal information can be used
to approximate a metric scale (Shepard 1980). And this is the crux of what
I have to say. I will claim that all perceived spaces are really ordinal.
However, sometimes these spaces can be said to converge on a metric
space.

Framing the Problem(s)


What I have to say is framed by the title of the conference “Reconceiving
Pictorial Space,” on which this volume is based, and on three questions
that seem to underlie its motivation. Each query comes in sequence as a
slight refinement of the previous.

11.2 IS PERCEIVING A PICTURE LIKE PERCEIVING THE WORLD?

Yes—and for some pictures to a large extent. This is one reason nonpro-
fessional, candid photographs work so well; the cinema can act as such
a culturally important surrogate for the everyday world; and precious
little experience, if any, is needed to appreciate the content of pictures
(Hochberg and Brooks 1962) or film (Messaris 1994). Nevertheless,
almost all theorists who bother to address this question—and few actu-
ally do—answer largely in the negative, choosing to focus instead on dif-
217 Reconceiving Perceptual Space

ferences between pictures and the world, regardless of how they define pic-
tures or they might define “the world.” Consider views from the humani-
ties and then from psychology.
Among artists and art historians, statements about the similarity of
pictures and the world are not prevalent. To be sure, few would deny
the impressiveness of certain trompe l’oeil (e.g., Cadiou and Gilou 1989;
Kubovy 1986) as having the power to be mistaken for a certain type of
reality, but equally few would claim this is other than a relatively small
genre, and it does not legitimately extend even to photographs. Moreover,
the effectiveness of trompe l’oeil is predicated, in part, on a tightly con-
strained range of depicted distances from the observer, or simply depths.
Rather than concentrating on similarity or verisimilitude of images and
their naturalistic counterparts, many in the humanities have focused on
viewer response. Responses to images are often indistinguishable from
responses to real objects, and this was an important problem in the
development of the Protestant Church in the sixteenth century and in the
Catholic Counter Reformation. Worship with images could not always be
separated from the worship of images, a violation of the Old Testament’s
First Commandment (Michalski 1993). This flirtation probably contri-
buted to the prohibition of images in Judaic and Islamic worship as well.
Over a broader cultural sweep, Freedberg offered powerful analyses of
how pictorial objects evoke the responses in people as real objects, but he
never claimed that pictures are mistaken for reality. To be sure, “people
are sexually aroused by pictures . . . ; they break pictures . . . ; they kiss
them, cry before them, and go on journeys to them” (1989, p. 1), but they
don’t actually mistake them for the real objects they represent. Instead,
images (and sculptures) stand in reference to the objects they represent,
attaining an equal status with them, to be loved, scorned, appreciated, or
decried in full value.
More generally, there are simply the difficulties of generating mim-
icry. Ernst Gombrich, for example, suggested that “the demand that the
painter should stick to appearances to the extent of trying to forget what
he merely ‘knew’ proved to be in flagrant conflict with actual practice. . . .
The phenomenal world eluded the painter’s grasp and he turned to other
pursuits” (1974, p. 163). Indeed, with the invention of photography there
was even a strong sentiment that, as Rodin suggested, “it is the artist who
is truthful and it is the photograph which lies. . . . [H]is [the artist’s] work
is certainly much less conventional than the scientific image” (Scharf 1968,
p. 226).
218 James E. Cutting

Within psychology, James J. Gibson used the picture versus real world
difference as a fulcrum to make a distinction between direct and indirect
(or mediated) perception. He is often quoted: “Direct perception is what
one gets from seeing Niagara Falls, say, as distinguished from seeing a
picture of it. The latter kind of perception is mediated” (1979, p. 147).
Although the nature and the wider ramifications of this distinction have
been much debated (e.g., see Cutting 1986a, 1998, for overviews), it is
clear throughout all discussions that picture perception and real-world
perception are conceived as different.1 Although few other psychologi-
cal theorists use pictures to discuss a direct/indirect distinction (but see
Sedgwick 2001), Alan Costall (1990), Margaret Hagen (1986), Julian
Hochberg (1962, 1978), John Kennedy (1974), Michael Kubovy (1986),
Sheena Rogers (1995), and John Willats (1997) all emphasize differences
between the world and pictures of it. Indeed, for William Ittelson (1996)
there are few, if any, similarities between the perception of pictures and
the everyday perception of reality.
In most psychological discussions of pictures versus reality, the essen-
tial element centers on the truism that pictures are two-dimensional sur-
faces and that the world around us is arrayed in three dimensions. At
their photographic best, pictures are frozen cross sections of optical arrays
whose elements do not change their adjacent positions when the viewer
moves. In particular, what is left of a given object seen in a picture from
a given position is always left of that object; what is right, always right;
and so forth.2 This is not always true in the natural environment. As one
moves forward, to the side, or up or down, objects in the world cross over
one another, changing their relative positions. In the world the projective
arrangement of objects is not frozen.
To be sure, there are other differences between pictures and the world
than the 2-D versus 3-D difference and the lack of motion in pictures. At
a comfortable viewing distance, the sizes of objects as projected to the eye
are generally smaller in pictures than in real life; pictures typically have a
compressed range of luminance values (and often of color) compared to
the real world; and there are lens effects that compress or dilate space. I
will focus on lens effects later, but let’s first consider a second question.

11.3 IS PERCEIVING PHOTOGRAPHIC SPACE LIKE PERCEIVING


ENVIRONMENTAL SPACE?

Again, yes—more or less—and, I claim, certainly more so than less so.


However, to answer this question, one must begin with an understanding
219 Reconceiving Perceptual Space

of the perceived space in the world around us. Variously, I will call this
environmental space, physical space, and even reality. My intent in this
multiplicity (other than to avoid semantic satiation) is to emphasize the
assumption that I consider all of these identical.
Two interrelated facts about the perception of physical space must be
considered. First, perceived space is anisotropic (Luneburg 1947; Indow
1991). In particular, perceived distances are somewhat foreshortened as
compared to physical space, particularly as physical distances increase.
This fact has been noted in various ways by many researchers (e.g., Gogel
1993; Loomis and Philbeck 1999; and Wagner 1985; see Sedgwick 1986,
for a review), although some methods of judging distance yield quite dif-
ferent results from others (Da Silva 1985; Loomis et al. 1996).
Second, this compression is likely due to the decrease in information
available. Such decreases in available information have been shown ex-
perimentally by Teodor Künnapas (1968) in the near range (/4 m) of
physical environments. They have also been demonstrated for near space
in computer-generated ones (Bruno and Cutting 1988). I contend that
the dearth of information about depth is responsible for compressions
throughout the visual range, particularly at extreme distances.

The Shape of Perceived Environmental Space


We don’t notice that our perception of space is not veridical. Where it is
compressed most, it is sufficiently far away that it has no consequence for
our everyday behavior; any “errors” in distance judgment would not usu-
ally matter.3 The amount of compression throughout perceived space is
often said to be well captured psychophysically by an exponent. That is,
perceived distance (w) is a function of physical distance (◊) with an expo-
nent less than 1.0, w; f (◊/1). To my knowledge, all previous investiga-
tions have assumed that the exponent is constant throughout perceived
space, and I think this is incorrect. Instead, it seems that exponents near
the observer are near 1.0, then generally decrease with the distance range
investigated. Shown in the top panel of figure 11.1 are plots for eleven
studies done in naturalistic environments and including distances beyond
20 m (see Da Silva 1985, for a more complete review). The horizontal
lines indicate the range of egocentric distances investigated (in meters);
their height in the figure corresponds to the value of the exponent best
fitting the data. Notice that exponents are generally from near 1.0 to
.65 for studies investigating ranges out to about 300 meters but nearer to
.40 for the only naturalistic study I know done beyond 1 km (Flückiger
1991).4
220 James E. Cutting

Figure 11.1
The top panel shows the exponents of perceived distance functions plotted by the
range of egocentric distances investigated in eleven studies. Notice the general
decline with increasing distance. This suggests an accelerated foreshortening of
space, from near to far. All studies chosen were done in naturalistic environments.
Those included, listed by order of decreasing exponents are Da Silva and Rozen-
straten (1979, reported in Da Silva, 1985), method of fractionation; Gårling (1970),
magnitude estimation; Harway (1963), method of partitioning; Teghtsoonian
(1973), magnitude estimation; Miskie et al. (1975), magnitude estimation; Da Silva
(1985), magnitude estimation; Kraft and Green (1989), verbal reports from pho-
tographs; Gilinsky (1951), method of fractionation; Flückiger (1991, two studies),
method of reproduction. The bottom panels select the data of naive viewers from
Gibson and Bergman (1954) and the data of Kraft and Green for special analysis.
That is, data are shown first with exponents and full egocentric depth range but
then with ranges incrementally truncated in the near range. This truncation sys-
tematically lowers the exponents. This means that the shape of perceived space
beyond about 50 to 100 m compresses faster than is captured by exponential
analysis. Exponents for the studies of Flückiger (1991), Gibson and Bergman
(1954), and Kraft and Green (1989) were calculated from the published data
for this analysis; others are available in Da Silva (1985).
221 Reconceiving Perceptual Space

With the addition of the data of Michelangelo Flückiger, the pattern in


the top panel of figure 11.1 looks as if there is a trend: as the range moves
outward from the observer, the exponent decreases. But is this just an
anomaly of an outlier study? A cleaner analysis of this effect can be seen
in the lower panels. Here, the data sets of Eleanor Gibson and Richard
Bergman (1954, for their naive subjects) and of Robert Kraft and Jeffrey
Green (1989) are considered. First, the complete published data sets were
fit with a series of exponents and a best-fitting exponent found. Next, the
data points nearest to the observer were omitted from the data sets
and the procedure for finding best exponents repeated. Truncation and
refitting procedures were repeated again until the data became unstable
or too sparse for further analysis.5 Notice that in both cases the truncation
of near data caused the best-fitting exponent to diminish. This can only
happen if perceptual space compresses faster than is captured by a single
exponent. Consider next the information available in the real world as the
beginnings of an explanation for such compression.

Information in Environmental Space


Peter Vishton and I (Cutting and Vishton 1995), among many others
(Landy et al. 1995; Luo and Kay 1992; Massaro and Cohen 1993;
Terzopoulos 1986), suggested that many sources of information about the
layout of the world around us are combined in complex ways. Cutting and
Vishton (1995) provided a list of nine different sources of information
used—occlusion, relative size, relative density, height in the visual field,
aerial perspective, binocular disparities, motion parallax, convergence, and
accommodation—although similar lists are stated or implied in works
from Euclid (Burton 1945) to Hermann von Helmholtz (1911/1925) and
continuously to the present day.6
These sources of information often measure the world in different ways.
For example, occlusion offers only ordinal information (what is in front
of what but not by how much); relative size has the potential of offering
ratio-scaled information (if two objects are identical in physical size and
one subtends half the visual extent of the other, then it is twice the dis-
tance from the observer); and stereopsis and motion perspective may
yield absolute distance information (e.g., Landy, Maloney, and Young
1991). Because it is not possible to compare the efficacy of measurement
on different scales, we employed scale convergence (Birnbaum 1983).
That is, we reduced the measures of information for each source to
their weakest shared scale—ordinal depth. We next located threshold
222 James E. Cutting

measures for ordinal depth in the literature and then used them to make
suprathreshold comparisons (Cutting and Vishton 1995). The results are
shown in the top panel of figure 11.2.
In the panel, threshold functions for pairwise ordinal distance judg-
ments are shown for nine sources of information known to contribute to
the perception of layout and depth. They are based on various assump-
tions and data and applied to a pedestrian. Inspired by and elaborated
from Shojiro Nagata (1991), the data are plotted as a function of the mean
distance of two objects from the observer (log transformed) and of their
depth contrast. Depth contrast is defined as the metric difference in the
distance of two objects from the observer divided by the mean of their
two depths. This measure is similar to that of Michelson contrast in the
domain of spatial frequency analysis.
Notice that, plotted in this manner, some sources are equally effica-
cious everywhere. In particular, I claim that potency of occlusion (the
most powerful source of information), relative size, and relative density do
not attenuate with the log of distance. Depth contrasts of 0.1%, 3%, and
10%, respectively, between two objects at any mean distance are sufficient
for observers to judge which of the two is in front. However, other sources
are differentially efficacious. Most—accommodation, convergence, binoc-
ular disparities, motion perspective, and height in the visual field—decline
with the log of distance. One, aerial perspective, actually increases with
the log of distance (it is constant with linear distance), but this source of
information, I claim, is used generally to support luminance contrast infor-
mation for occlusion. Notice further that, integrated across all sources,
the “amount” and even the “quality” of information generally declines
with the log of distance. This seems the likely cause of compressed per-
ceived distances.
The general shapes of the functions in figure 11.2, plus a few practical
considerations, encouraged us to parse egocentric space into three regions
(see also Grüsser and Landis 1991). Personal space, that which extends a
little beyond arm’s reach, is supported by many sources of information
and is perceived almost metrically (Loomis et al. 1996), that is, with an
exponent of 1.0. This means that observers can generally match a lateral
distance of, say, 40 cm with a 40 cm distance extended in depth. After 2 m
or so, height in the visual field becomes an important source (a normal
adult pedestrian cannot use this information until objects are at least
1.5 m distant), and motion perspective for a pedestrian is no longer a blur;
but accommodation and convergence cease to be effective.
223 Reconceiving Perceptual Space

Figure 11.2
The upper panel shows the ordinal depth thresholds for nine sources of infor-
mation (panel modified from Cutting and Vishton 1995). Egocentric depth (the
abscissa) is logarithmically represented. Depth contrast (the ordinate) is measured
in a manner similar to Michelson contrast in the spatial frequency domain. That
is, the difference in the metric depths of two objects under consideration (d1 : d2)
is divided by the mean egocentric distance of those objects from an observer
[(d1 + d2)/2]. Egocentric space is also segmented into three general regions, which
grade into one another. Personal space, out to about 2 m, is perceived as Euclidean
(exponent of 1.0); action space, from about 2 to about 30 m, demonstrates some
foreshortening (exponents near 1.0); and vista space, beyond about 30 m, often
demonstrates considerable foreshortening (exponents declining to .40 or so, as
suggested in figure 11.1). Different sources of information seem to work differen-
tially in the different regions. An assumption made is that threshold measure-
ments, shown in the panel, are good indicators of suprathreshold potency. The
lower panel shows those thresholds isolated for pictorial sources of information,
with personal space excluded because few pictures (before the twentieth century)
depict space within 2 m of the viewer.
224 James E. Cutting

We suggested that at this distance (at about 1.5–2 m) there is a bound-


ary, and it marks the beginning of action space. This new space extends
out to about 30 m or so. Using 10% depth contrast as a general threshold
of utility, both disparities and motion perspective cease to be useful at this
distance. Most practically, an average individual can throw something
reasonably accurately to a distance of 30 m. But perhaps most interest-
ingly, 30 m is about the height of the Andrea Pozzo ceiling in the Chiesa
di Sant’Ignazio in Rome (1694), where pillars and architectural ornaments
are painted on a vaulted ceiling so that real pillars and painted pillars are
indistinguishable when observed from a designated viewpoint (see Pirenne
1970). Action space is perceived near-metrically; its exponents are often
only slightly less than 1.0.
Finally, there is vista space, everything beyond about 30 m. Here only
the traditional pictorial sources of information are efficacious, and vista
space becomes increasingly compressed with distance. Exponents in the
near vista may be near 1.0, but those near the horizon become distinc-
tively less.
Two caveats: First, although the boundaries between the spaces as we
have drawn them are not arbitrary, neither are they well demarcated.
Anywhere between 1 and 3 m could serve as the personal/action boundary
and anywhere between 20 and 40 m could serve as the action/vista bound-
ary. Training, stature, and a variety of other variables could also influence
these boundaries. Most emphatically, one space should be expected to
grade gracefully into the next.
Second, some functions can change (Cutting and Vishton 1995). If, for
example, one places an observer in a British sports car driving through the
countryside, then the function for height in the visual field would move a
bit to the right. This is due to adjusting to the reduction in eye-height to
about 1 m and because motion perspective would move considerably to
the right (adjusting to an order of magnitude change in optical velocity).
Some other functions change with the environment—height in the visual
field as plotted is for vision standing on a plane (hills introduce changes in
slant, altering the function, and occluded regions that create discontinu-
ities in the function); aerial perspective is plotted for a clear day (humid-
ity, mist, and fog would move the function considerably to the right). Still
other functions may not prove useful in a given setting—there may be no
objects on which to compare relative sizes. Finally, some functions change
with characteristics of the observer: some people are stereoweak, others
stereoblind, and sometimes this can occur after only a few hours of stereo
225 Reconceiving Perceptual Space

deprivation (Wallach, Moore, and Davidson 1963); many people over the
age of forty can no longer accommodate; and, of course, adults are of dif-
ferent sizes and youngsters crawl and toddle, further changing the func-
tion for height in the visual field. All of these factors contribute, I claim,
to the lability of perceived distances across people and environments.

Information in Pictorial Space


Most relevant to the purposes of this volume, I will claim that parts of this
scheme can be applied directly to the perception of most photographs and
to some art (particularly that using linear perspective). Again, such an
application is suggested in the lower panel of figure 11.2, where pictorial
sources of information—occlusion, relative size, relative density, height in
the visual field, and aerial perspective—are shown. The first four comprise
what was traditionally called artistic perspective (and with the addition of
receding parallels, linear perspective; Kubovy 1986), with the fifth being
added by Leonardo da Vinci (Richter 1883). In addition, at least with
respect to art before the twentieth century, few paintings depict anything
within personal space. Portraiture and still life—certainly the most proxi-
mal of artistic genres—typically portray their objects from just beyond
arm’s length, if not a greater distance. Moreover, this generally remained
so even after the invention of photography. Thus, in the lower panel of
figure 11.2, personal space is omitted.
Of course, there is additional information in pictures—information
specifying that there is no depth (see discussion of flatness in Rogers,
chapter 13 in this volume; Sedgwick, chapter 3 in this volume). When
holding a photograph, the status of our accommodation and conver-
gence, and the lack of binocular disparities and of motion perspective, all
tell us that we are not really looking into the distance. Thus, pictures are
endemic situations of information conflict. I think we fairly readily dis-
count such conflict, although there may be residual influences. Reducing
such potential conflict is one reason cinema is typically viewed from a
distance, often 20 m or so, where accommodation and convergence are
inactive and binocular disparities not salient.

A Case Study with Different Lenses


Representations of space often have manifold consequences; they even
cause a perceptual elasticity. This changeableness can be demonstrated in
at least two ways—in the effects of lenses (e.g., Farber and Rosinski 1978;
Lumsden 1980; Pirenne 1970), and in the effects of viewing pictures from
226 James E. Cutting

noncanonical viewpoints of pictures, such as to the side (Cutting 1986b,


1987). Only the former will be discussed here.
Consider how camera lenses transform information for depth. Figure
11.3 shows two images from Charles Swedlund (1981) with the same con-
tent but taken with different lenses—a 35 mm lens (a somewhat short lens
for 35 mm film) and a 200 mm lens (a telephoto). These images have been
cropped, their sizes have been adjusted, and they were not taken from the
same viewpoint. Instead, Swedlund normalized the relative sizes of the
sign at the right of both images so that the sign takes up roughly the same
proportion of the image space. To do this Swedlund was about four times
farther away from the sign in the lower image, as noted in the plan view
in the right panel of the figure. But look at the basketball backboards to
the left in the background. These have grown enormously in size in the
lower image, and their difference in relative size has diminished. The
latter is due to the compressive effect of telephoto lenses. Ordinarily, a
200 mm lens blows up space 5.7 times compared to a 35 mm lens. With
the 5.7-fold increase in image size, there should be a 5.7-fold decrease
in apparent image depth. This is not quite true in the Swedlund (1981)
images, the deviation due to cropping and sizing.
How exactly does a telephoto lens compress depth? Telephoto lenses
enlarge the image for the observer. All enlargements compress depth—
those created by longer lenses and those created by moving the observer
forward toward the image from the original station point (the point of
composition, or the point from which, when looking at the image, the
optics are reconstructed most perfectly). How compression occurs is
schematically suggested in figure 11.4. Because the points on the image
do not change, and because the lines of sight from the observer to these
points can be extended into “pictorial space,” they find their end points
at different pictorial depths. The closer the observer moves toward an
image, the more compression in depth there should be; the farther he
or she moves away from the image, the more dilation there should be.
Sometimes these effects are demonstrable in experiments, but sometimes
they are not. One point I would like to emphasize is that our perception
of spaces is somewhat elastic; sometimes we pick up on information,
sometimes not.
But what information is changed with a long lens? This is a bit more
difficult to explain. Enlargements play with familiar size. A person 10 m
away would ordinarily be in action space; a person 100 m away is ordi-
narily in vista space. Seen through a 500 mm lens, however, the latter
227 Reconceiving Perceptual Space

Figure 11.3
The two images on the left are from Swedlund (1981), reprinted with permission
of Holt, Rinehart, and Winston. These images show the changes in pictorial space
due to changes between a short and a long lens. The point from which the bottom
image was photographed was about four times the distance of that for the top
image, and both were cropped and sizes adjusted to make the appearance of the
sign in the foreground the same. The diagram to the right is a plan view of the lay-
out of the sign, the two backboards, and the two camera positions. It was recon-
structed knowing that backboards are usually about 25 m apart and from the
measurements of their heights (in pixels) in the digitized images. (Photographs
from Photography: A handbook of history, materials, and processes, 2nd edition by
Charles Swedlund and Elizabeth U. Swedlund, copyright © 1981 by Holt, Rine-
hart and Winston, reproduced by permission of the publisher.)
228 James E. Cutting

Figure 11.4
An affine reconstruction of pictorial space. The left panels show slices through
untransformed pictorial space as seen from the composition point and the image
seen by the observer; the other panels show transformations and views due to
changes in observer position away from the composition point. Top panels show
the affine transform in horizontal planes; the bottom-right panel shows a similar-
ity transform (enlargement). After Cutting (1986b).

person has the same image size as the former seen through a 50 mm
lens. Thus, familiarity with sizes of objects in the environment shifts, in
this case, the whole of the environment shifts one log unit toward the
observer; familiarity with what one can do with those objects move distant
ones into what appears to be action space. (In telephoto images, as in reg-
ular images, there typically is no content in personal space.) Thus, what
passes for action space now really extends out to perhaps 300 m.
If we carry with us a set of expectations about objects and the efficacy
of information in real space to what we see in the telephoto image, per-
ceptual anomalies will occur (Cutting 1997). Compared to the perception
of environmental and normal pictorial spaces, relative size differences and
relative density differences are reduced. Consider size. Basketball back-
boards are normally about 25 m apart. In the Swedlund images in figure
11.3 the ratio of sizes (linear extent) of the two backboards is 1.5:1 in the
top image but only 1.24:1 in the bottom image. The perceived distance
229 Reconceiving Perceptual Space

difference in the top image could rather easily be 25 m; that in the lower
image could not. In near space, we expect smaller relative size differences
for similar objects closer together. In addition, height in the visual field
also changes; the difference in height between the base of the poles for the
two backboards is three times greater in the telephoto image. We expect
larger height-in-the-visual-field differences for nearer objects.

The Shapes of Perceived Photographic Space


Consider next the data set of Kraft and Green (1989), already consid-
ered earlier as if it represented judgments about depth in the real world.
Although gathered for different purposes,7 these data are a definitive
example of lens compression and dilation effects. Kraft and Green pre-
sented many photographs to observers. These were taken with a 35 mm
camera and five different lenses: 17, 28, 48, 80, and 135 mm (where the first
is called a short lens and the last a telephoto lens). With such lenses the
horizontal field of view subtends about 105, 60, 45, 32, and 20°, respec-
tively. Because a 48 mm lens is essentially a standard lens for a 35 mm
camera (35 mm is the measure along the longer edge of the film frame;
about 50 mm is the measure along its diagonal, and the functional diam-
eter of the image circle; see, for example, Swedlund 1981), the layout of
objects and their depth should appear relatively normal. Moreover, since
the time of Leonardo, recommendations have been made to keep the
larger, usually horizontal, subtense of a picture within 45°–60° to avoid
distortions (Carlbom and Paciorek 1978). In two different outdoor envi-
ronments, Kraft and Green (1989) planted poles at distances of 20, 40, 80,
160, and 320 m from a fixed camera. Viewing different arrangements of
fifty slides, observers made judgments of the distance of each pole from
the camera. Mean results are shown in the left panel of figure 11.5. This
fan-shaped pattern provides grist for further analyses.
Notice first that perceived distances are generally quite a bit less than
real distances, even for those stimuli shot with the standard lens. Ideally
this should not occur.8 To assess this general effects of foreshortening due
to the experimental manipulation (the photographic lenses) and the ex-
perimental situation, I regressed the mean perceived distances in the left
panel of figure 11.5 against a parameter cluster: ◊ · (50/L). This is the
physical distance of the posts from the camera multiplied by 50 (the
standard lens length) over the lens length used. The second half of this
composite variable is the extent to which perceived space ought to be
compressed (50/L/1) or dilated (50/L01; Cutting 1988; Lumsden 1980).
230 James E. Cutting

Figure 11.5
The data of Kraft and Green (1989) are presented in the left panel, showing per-
ceived distances of posts as functions of their physical distance from a camera and
its lens. The middle panel shows a model prediction based on the combination of
physical distance and lens length; and the right panel shows a prediction based on
this parameter and an exponent of .68. The fit shown in the left panel is statisti-
cally superior to that in the middle panel, revealing compression in perceived
depth.

Thus, for example, a 135 mm lens should compress depth by a factor of


about 2.8 (multiplying depth values by 0.37); a 17 mm lens should com-
press it by a 0.35 (actually dilating perceived space, multiplying values by
2.85).
This analysis accounted for 96% of the variance in the data (R2;.96,
F(1, 23);595, p / .0001), fitting the pattern of data extremely well. The
weight (b) given the parameter g · 50/L in the equation was .47, about half
what might have been expected. Perceived half-depth could be a compro-
mise between full depth (carried by height in the visual field, relative size,
relative density) and zero depth (enforced by no disparities and no motion
perspective). Regardless, the schematized data, however, can serve as post
hoc “predictions” for (i.e., model fits to) the real data. These are shown in
the middle panel of figure 11.5. The match between them is quite good but
not quite good enough.
Notice that there are some distinguishing characteristics in the shape of
the predicted functions that are not generally wanted. Most salient is that
231 Reconceiving Perceptual Space

at large physical distances (g) there is insufficient foreshortening in the


perceived distances (w) when compared to the Kraft and Green data. This
means that beyond general compression, there is general change in com-
pression rates in the data, suggesting exponents less than 1.0. I reran the
regression analysis with various exponents to determine the value of the
best fit. A single exponent of .68 captured the most variance, increas-
ing that accounted for to 99.1% (F(1, 23);2660, p / .0001), a reliable
improvement ( d 2;7.2, p/ .01) over the exponentless prediction. The
model fits with the exponent are seen in the right panel of figure 11.5.
(Remember, truncating the data at the near end decreases the exponent,
as shown in the lower panels of figure 11.1.)
Of course, it may be simply a coincidence that judgments of distance in
photographs (Kraft and Green 1989) and judgments in the real world
(e.g., Da Silva 1985; Wagner 1985) are so similar, as shown in the top
panel of figure 11.1. In may also be that the similarity in the decline in
exponents found by truncating the near-distance data is similar in both
photographic data (Kraft and Green 1989) and in real-world data (Gibson
and Bergman 1954), shown in the lower panels of figure 11.1, is also a
coincidence. But I think not. Most pleasantly, there are some elegant data
gathered at ZiF (Center for Interdisciplinary Research), the site of the
conference on which this volume is based. These show similar effects in the
real world and in photographs of it (Hecht, van Doorn, and Koenderink
1999).9 These data, if not the reanalyses that I have presented here, ought
to be sufficiently convincing of the strong similarity—indeed functional
identity—between photographic and real-world spaces.

11.4 IS PERCEIVING THE LAYOUT IN OTHER PICTURES LIKE


PERCEIVING THAT IN ENVIRONMENTAL SPACE?

Yes—at least in terms of principles if more than the end result. Earlier I
had suggested that psychologists and others interested in pictorial space
have spent perhaps too much time with photographs and linear perspec-
tive paintings, drawings, and etchings. These are among the more recent
of pictures that human beings have crafted. Manfredo Massironi and I
(Cutting and Massironi 1998) suggested one might equally start at the
other end of time, at least from the perspective of human culture. Cave
paintings—some of which are at least 30,000 years old—as well as car-
toons, caricatures, and doodles are made up of lines. These pictures are
not copies of any optical array, and yet as images they depict objects well.
232 James E. Cutting

Many of these compositions can be read to represent animals in depth,


often by occlusion of one by another.
How do line drawings work? What kind of space do they build? Cutting
and Massironi (1998) suggested that there are four basic types of lines and
that these ordinally parse the regions on either side of the line or parse
themselves from those regions. The four types are shown in figure 11.6.
Edge lines separate a figure from background, where one side belongs to
the figure and is closer to the viewer and the other side belongs to the
background. We notate this type of configuration [aAb] or [bAa], where
letter order denotes ordinal position in depth, [a] in front of [b]; and
uppercase [A] denotes the line, lowercase [a] one of the two regions on
either side of the line. A second kind is an object line, where the line stands
for an entire object in front of the background [bAb]. Its inverse is the
crack line [aBa], where the line invites the viewer to imagine the interior
space hidden from view. Finally, there are texture lines, which subdivide
again. Texture lines can represent small edges, small objects, or small
cracks; and they can also represent shading, color, or what the tradi-
tional pedagogical literature on drawing calls mass (Speed 1913).
Consider an example from the oldest known works of paleolithic art.
Shown in figure 11.7 are several rhinoceroses from La Grotte Chauvet in
the south of France (see also Bahn and Vertut 1997; Chauvet, Brunel
Deschamps, and Hillaire 1995; Clottes 2001). As part of this figure one
can see edge lines (outlining a rhinoceros), object lines (its horns), crack
lines (its mouth), as well as texture lines (the coloring of the flank). The
antiquity of this image, and the fact that it has the complete set of line
primitives outlined by Cutting and Massironi (1998), suggests that line
primitives are not culturally limited or even culturally determined. The
fourfold use of lines seems deeply embedded in our genome.
I claim that each line serves to carve up a small bit of pictorial space;
each contributes a local bit of ordinality. Nothing is needed in this case
but a segregation of two different depth planes—one of the animal and
one of the background against which it stands. As images get richer in
information, depth can become richer as well. Consider another example
from La Grotte Chauvet in figure 11.8.
Here a few well-crafted lines represent four horses. There is great con-
troversy about whether or not these horses should be read as occupying
the same space, each occluding another that is higher up in the pictorial
place. Nonetheless, I think depth is inescapable. In the paleolithic corpus
of images, one often finds the superimposition of lines from different
233 Reconceiving Perceptual Space

Figure 11.6
A taxonomy of lines. Edge lines separate the regions on either side and assign dif-
ferent ordinal depth. Four figures that play with this relationship are shown: (a)
the faces/goblet illusion, after Rubin (1915); (b) an ambiguous occlusion figure,
after Ratoosh (1949); (c) the devil’s pitchfork, after Schuster (1964); and (d) a rec-
tangle/window, after Koffka (1935). Next are shown some object lines: (e) is a mid-
twentieth-century version of a television antenna; and (f) shows the twigs at the
end of a tree branch. Third are figures with crack lines: (g) is the mouth of a clam,
after Kennedy (1974); and (h) is a crack in a block. Note that (f) and (h) are exact
reciprocals—switching from object line to crack line. Finally, four texture line
types are shown: (i) texture edges of occlusion in cobblestones, after de Margerie
(1994); (j) texture objects as palm fronds, after Steinberg (1966); (k) texture cracks
as mortar between bricks, after Brodatz (1966); and (l) texture color, indicating
shadow. (Adapted from Cutting and Massironi, 1998)
234 James E. Cutting

Figure 11.7
A thresholded detail from a painting in La Grotte Chauvet demonstrating the use
of four types of lines—edges, objects, cracks, and textures. Because this image is
at least 30,000 years old, it would appear that this typology of lines is not cultur-
ally relative but possibly biologically engrained. (Image reprinted from Clottes,
2001, with the kind permission of Jean Clottes, Ministère de la Culture)

animals in different orientations, lines crossing, suggesting that they were


composed without respect for one other. Here, however, we find what
appear to be true occlusions of horses each in the same orientation. Each
edge line can be interpreted locally to generate depth, and local depth
assignments can be written [bAa]. However, taken as an ensemble, sev-
eral relations have to be rewritten, pushing some regions and lines farther
into depth. Thus, the occluding edge of the second horse becomes [cBb],
the third [dCc], and the fourth [eDd]. In this manner global coherence can
be built up from local depth assignments, and multiple, ordered depth
planes can be built up as well. I suggest this is done in a manner not much
different than seen in cooperative stereo algorithms (Julesz 1971; Marr
and Poggio 1976), although in pictorial cases it is much simpler. Depth is
incrementally built up, in more difficult cases perhaps with many passes
through the data. Moreover, there can be great play in this. Figure
11.6a–6d are examples from the psychological literature, but of course
235 Reconceiving Perceptual Space

Figure 11.8
From pairwise ordinal depth relations to five ordinal depth planes. A thresholded
detail from perhaps the most famous and controversial image in La Grotte
Chauvet. Many claim it should not be read as four horses in depth, although I dis-
agree. If it is read in depth, one can follow the assignments of occlusions, using
edge lines [bAa]. Each of four pairs has this assignment. If this information is fed
into a comparator that knows that objects can occlude, the end result is five ordi-
nal pictorial depths, [a]–[e]. (Image reprinted from Clottes, 2001, with the kind
permission of Jean Clottes, Ministère de la Culture)

much of the corpus of M. C. Escher (Bool et al. 1981) can be construed


this way.
Assignments of edge lines [aAb] and [bAa] are often judgments of occlu-
sion. Assignments of many lines (and all line types), however, could also
be made on the basis of height in the visual field; assignments of similarly
shaped objects (made up of lines) could also be made on the basis of rela-
tive size; and so forth. Thus, bringing knowledge to bear on a graphic sit-
uation, one can assign depth order to many elements, The five pictorial
sources of information—occlusion, height in the visual field, relative size,
236 James E. Cutting

relative density, and aerial perspective—can constrain one another quite


well so that a reasonably rich, affine space is established. Moreover, an
approximation to a metric space can also be achieved, particularly with
the application of linear perspective.
How might this be done? The analogy I wish to make is with respect to
nonmetric multidimensional scaling (Kruskal 1964; Shepard 1980), and it
is a deep lesson in measurement theory. By considering only pairwise ordi-
nal relations among all (or even only most) members of a set of data, one
can assemble a nearly metric space of that set. Metric distances, even
ratios, are not needed. Yacov Hel-Or and Shimon Edelman (1994) demon-
strated the logical efficacy of this approach to qualitative stereo, and
Cutting and Vishton (1995) suggested it should work considering depth
in general. Thus, one can think of perceived space—even at its articu-
lated, near-Euclidean best—as built up incrementally from constraints of
ordinality. These constraints, when sufficiently rich, converge on a near-
Euclidean framework.
In summary, I claim that perceiving depth in pictures and perceiving
depth in the real world are cut from the same informational cloth. Thus,
if we misunderstand perceived pictorial space (as might be suggested by
the title of the conference on which this volume was based), then it follows
that we misunderstand perceived environmental space as well. I claim
further that when ordinal depth information is sparse, perceived depth is
also crude, confined to a few depth planes. When ordinal information is
richer, perceived space becomes more articulated, allowing first for many
depth planes (and an essentially affine representation). When that infor-
mation is extremely rich, as with the addition of binocular disparities and
motion perspective in cluttered environments and in daylight, ordinal con-
straints can become sufficiently tight to approach a metric representation.
Thus, the concatenation of ideas from measurement theory and from
the computational practice of multidimensional scaling allows promise of
considering the perception of space in pictures and in the world as pro-
ceeding from the same principles.

ACKNOWLEDGMENT

I thank Claudia Lazzaro for many fruitful discussions and for her patience
in listening to many ill-formed ideas, Robert Kraft for sharing method-
ological information about this study of pictorial depth, and Patrick
Maynard for some insightful comments.
237 Reconceiving Perceptual Space

Notes
1. Earlier in his career, however, Gibson was less sure: “It is theoretically possible
to construct a dense sheaf of light rays to a certain point in a gallery or a labora-
tory, one identical in all respects to another dense sheaf of light rays to a unique
station point thousands of miles away” (1960, p. 223), although he denied it was
obtainable in practice.
2. This is an axiom from Euclid’s optics (Burton 1945). Euclid’s axioms and his
proofs deal with both stationary and moving observers.
3. This is not to say that people cannot be trained to perceive vista space more
veridically; clearly they can, and the data of E. Gibson and Bergman (1954)
and Galanter and Galanter (1973) show this. The focus of this chapter, however,
is on what normal adults, without specific distance training, will perceive and
judge.
4. Flückiger (1991) had observers judge the distance of boats floating on Lake
Leman, looking from near Geneva toward Montreux. Only relative size and
height in the visual field would provide firm information, perhaps with the addi-
tion of aerial perspective and relative densities of waves. Experimental results for
judgments between 0.2 km and 2.0 km and between 0.75 km and 2.25 km were
quite tidy; whereas those between 2.8 km and 5.6 km were not. Flückiger did not
fit exponents to his data, but this is easily done. Mean exponents for the first com-
plete data sets were .36 and .44, respectively. In addition, I have omitted the data
of Galanter and Galanter (1973) here. Their observers were highly trained and,
although they viewed objects up to 10 km into the distance, they viewed them
from airplanes at altitudes of about 60 m. This raises the function of height in the
visual field by a factor of 40, as discussed in the next section.
5. Exponents become unstable when a first pair exhibits overconstancy (more per-
ceived distance between them than physically present). When this occurs, expo-
nents tend toward zero. This also happened in the analysis of the data of Flückiger
(1991).
6. Linear perspective, often cited as a single source of information, is really a con-
comitance of occlusion, relative size, relative density, and height in the visual field,
using the technology of parallel straight lines and their recession.
7. Kraft and Green (1989) believed that the foreshortening of space seen in pho-
tographs due to the use of lenses of different lengths is caused by the truncation
of the foreground. Although this is a possibility, the purist form of the argument
is that the foreground is truncated but the remainder of space remains the same.
Their own data (Kraft and Green, 1989, experiment 1) are not consistent with this
idea; they show a fan effect rather than parallel lines.
8. The Kraft and Green article did not give full details of the presentation of the
stimuli. Thus, I contacted Robert Kraft and he kindly provided additional exper-
imental details (Kraft, personal communication, May 24, 2000). First, the length
of the lens used to project the images was 100 mm. This means that the images are
half as large than they would be with a 50 mm projector lens from the distance of
238 James E. Cutting

the projector. Ideally, when seen from the projector, such diminution should dilate
depth by a factor of about 2.0 (multiplying distances by about 2.0; see Cutting,
1988, for a similar analysis). However, the distance of the projector to the screen
was at 6 m, twice the mean distance of the observers (3 m). This latter effect
enlarges the image by a factor of 2, compressing depth (multiplied by 0.5). These
two effects are linear and should cancel. Theoretically one would expect perceived
distances to be compressed by a factor of 0.5 · 2.0 (multiplied by a factor of 1.0).
This also means that the images seen in the experimental situations subtended the
same angle as the corresponding images in the real world.
9. The most convincing comparisons in the data of Hecht, van Doorn, and
Koenderink (1999) are for angular judgments of the intersections of wall of build-
ings at ZiF, which has few right angles.
Chapter 12
Pictorial Space

Jan J. Koenderink and Andrea J. van Doorn

12.1 WHAT IS “PICTORIAL RELIEF”?

When you are handed a “straight” photograph of, say, a statue you can
either
• look at it and see a flat piece of paper covered with pigments in a certain
simultaneous order, or
• you can look into it and see an object (“the statue”) in “pictorial space.”

The former perception is of a “physical object” that has an objective


existence in the usual sense; the latter is a purely “mental entity.”
Pictorial space has no objective existence in the usual sense. It has noth-
ing per se to do with the fact that the photograph might have a certain his-
tory. For suppose that it—the physical object—turned out after all to be
a piece of paper covered with darkish fungus growth, such that you, erro-
neously as it turned out, had mistaken it for a photograph. Such facts
would be quite irrelevant with respect to the pictorial space. Indeed, it
is well known that people often see pictorial spaces in clouds or dirty old
walls (Leonardo da Vinci 1989). Thus “pictorial space” is a somewhat
mysterious entity; small wonder that little objective knowledge concerning
its properties has become established so far.
The phenomenology of pictorial space is familiar enough to people from
cultured societies. You experience the visual field—a two-dimensional
simultaneous presence—augmented with a “depth” dimension. The depth
is like a “flow away from you” (Hildebrand 1893/1945) that may be
checked or stopped by more or less “opaque objects.” When we say
opaque objects, we mean that you don’t see “through” them. This has
nothing to do with optics proper of course. We are outside the realm of
physics.
240 Jan J. Koenderink and Andrea J. van Doorn

Opaque objects manifest themselves through their “boundary,” a “sur-


face,” often called pictorial relief. A definitive reference on this is by the
German sculptor Adolf Hildebrand (1893/1945). As a sculptor Hildebrand
was mainly interested in actual objects, that is to say, artfully formed
chunks of marble, and not so much pictures of objects. Here we meet
with an unexpected twist, though, for Hildebrand essentially didn’t dis-
tinguish between pictorial space and the optical perception of physical
space in front of the observer! When “motion-related perceptions” are
excluded “optical space,” that is, what you see in front of you in a real
environment, and pictorial space are phenomenologically similar indeed.
Examples of motion-related perceptions are those associated with eye
movements (vergence), accommodation, binocular disparity, motion par-
allax, and so forth. The ontologies of optical space and pictorial space
may be taken to be quite dissimilar or rather similar depending upon your
point of view.
What inhabits pictorial space is usually described in terms of geometri-
cal or physical properties, or both, that refer to the generic structure of the
observer’s biotope. In fact, observers don’t seem to have any other option.
Thus you describe the face you see in a cloud in terms of properties of
actual human facial structures with which you are familiar, although you
know perfectly well that clouds are differently structured from faces and
you are convinced that there isn’t any face in the cloud from the perspec-
tive of the meteorologist, that is to say, that the face is a mere “illusion.”
By illusion, you mean a mental thing without objective reference. This
applies equally to pictorial and to optical space. A pair of hills may look
like “female breasts” (such as in the movie King Solomon’s Mines) and the
rippled water on a lake in the moonshine “silvery,” although there are no
giant women nor precious metals in the real landscape in front of you or
in the picture of the landscape that you may happen to behold. Nor need
you believe there are. Of course the real landscape looks real enough to
you, whereas the picture looks like a (real) picture of a possible landscape
to you.
A valid way to describe pictorial space in different terms is the way
of the artist. Artistic renderings can be absolutely compelling in many or
a few respects, although you are never “fooled,” because you know quite
well that you are looking at a certain rendering of a scene, not at a
real scene. Artistic statements are different in (objective) structure from
straight photographs, though. So far science has not been able to exploit
the uncanny ability of artists to produce such responses in human ob-
241 Pictorial Space

servers. Although this matter is highly interesting, also from a funda-


mental scientific perspective, we won’t pursue it here.
Observers bring expectations of a similar kind to both optical and pic-
torial space. Thus you may notice spontaneously that painted grapes look
“glassy” (or vitreous) and thus somehow not “right.” They don’t look to
you like real grapes “should” in either optical or pictorial space. Yet the
“grapes” in pictorial space may look like perfectly good pictures of grapes,
in a particular painterly style say. Glassy-looking grapes in optical space
don’t look right in any sense, and you may well suspect to be confronted
with hard to swallow (fake) specimens.
The scientific studies of optical and pictorial space may largely coincide
or be totally distinct according to whether you study the relation of per-
ceptions to the distal objects or study relations intrinsic to optical or pic-
torial space. In this chapter, the emphasis will be almost singularly on the
latter approach.

Issues of “Information”
Here is a remarkable fact: When you look at the scene in front of you
you pretty much see the scene in front of you. We mean this in the sense
that you find no problem in making judgments of either a geometrical
or a physical nature that you know are likely to be corroborated by vari-
ous explorations you might make were you to walk up to the scene and
explore it in various ways. Examples of geometrical attributes would be
sizes, distances, and spatial orientations; examples of physical attributes
would be such properties as being rough, smooth, metallic, wooden, or wet.
Although you know that your judgments are unlikely to be correct in
every detail, you are convinced that they are good enough for most inter-
actions with your trusty environment. When you get out of your generic
biotope, things may turn out to be different than expected, though, and
you have to be more careful; that is, you have to reckon with various
forms of erroneous—in the sense of leading you to unsuccessful action—
judgments. For instance, you are likely to seriously misjudge distances
and shapes when transferred to the submarine environment (Lythgoe
1979), and you will probably misinterpret X-ray pictures.
Because your only physical interaction with the scene is of an optical
nature, it must be the case that you interpret the “radiance” at the loca-
tion of your eye in terms of “information” relating to the geometry and
physical properties of the scene. For the moment we consider only monoc-
ular vision.
242 Jan J. Koenderink and Andrea J. van Doorn

The radiance is a function of position and direction that for each posi-
tion and direction specifies the photon spectral number density per unit
solid angle, unit area, unit time, and unit photon energy interval. It is sub-
ject to few constraints. Photon number density is nonnegative throughout,
and, at least in empty space, the radiance at different locations in the direc-
tion of their connecting line must be the same. Otherwise, the radiance is
quite arbitrary. The best intuitive way to think of radiance is as an infinite
storage cabinet containing all possible photographs, taken from any con-
ceivable viewpoint, field of view, view direction, spectral filter, and so
forth. Its structure is extremely complicated in the sense that when you
increase the resolution you obtain more and more variation. You would
see all the hairs on the body, the tiny organisms that live in the crevices of
the skin, all the tiny articulations of these, and so forth. In that sense the
radiance is not a simple function (i.e., it is not continuous and differen-
tiable) at all, but what mathematicians call a “distribution.”
From the perspective of physics the radiance is due to the scattering
of photons emanating from certain “primary sources,” for example, the
sun, by the scene to your eye. Multiple scattering is the rule. A photon
that arrives at your eye is likely to have been scattered many times by
the atmosphere and various objects in the scene. All photons of a given
energy are the same; they don’t carry an imprint of their history. It is only
the radiance as a whole that may be interpreted in terms of a history of
scattering events in the scene (Born and Wolf 1959).
Your eye transducts the radiance at the position of the cornea into a
simultaneous-successive order of action potentials in the optic nerve. The
retina doesn’t distinguish between the energies of photons once they have
been absorbed (see the “Law of Univariance”; as described by Wyszecki
and Stiles 1967), and many different patterns of photon absorptions lead
to identical optic nerve patterns. The structure of the “input to the brain”
reflects—in an extraordinarily complex way—the structure of the scene in
front of you (Zeki 1993). Some correlations between certain aspects of this
structure and certain aspects of the structure of the scene are understood,
or even well understood, but much still awaits discovery. That is one
reason, though not the most important one, why we deem it prudent to
distinguish between “structure” and “information.”
Structure simply refers to aspects of spatiotemporal order. It implies a
topology (adjacency in space-time) but has nothing to do with any refer-
ence to the scene. Information is structure referred to a real or imagi-
nary scene by an observer. Thus the concept of information implies an
243 Pictorial Space

observer with experience with generic scenes who posits a particular scene
and takes the available structure as referring to that scene. This is Franz
von Brentano’s “intentionality” (Intentionalität), which turns structure into
information in the sense of “meaning” (Brentano 1874). To take structure
for information in this way is an act of faith on the part of the observer.
Structure is not intrinsically information in any way, although most text-
books are designed to mislead you on the issue. That is why information
can be wrong, even information concerning the (real!) scene in front of
you. Structure is never wrong, nor is it ever true; it simply exists. Of course
the act of faith on the part of the observer need not be a conscious one.
It typically is not. Organisms simply have to stick their necks out or they
won’t survive.
Classically (Berkeley 1709/1975) optical information is discussed in
terms of “cues.” The meaning of cue is differently construed by various
authors. Here we take it to refer to structure interpreted by an observer in
terms of knowledge of “ecological optics” (Gibson 1950). “Knowledge” is
understood in the sense of “know-how,” or capability. Such knowledge is
of two kinds, generic and particular. Knowledge of generic ecological
optics coincides with the laws of physics as they apply to perception. An
example would be that large and small images of similar things indicate
near and far distances of these things. Generic knowledge holds good in
virtually all biotopes. In contradistinction, particular knowledge applies
to particular biotopes and may easily fail in the wrong setting. An exam-
ple would be that strong and otherwise unaccountable contrasts (“glitter”)
indicate wet or metallic surfaces. This won’t work in an art gallery where
chrome objects are rendered in paint on canvas. Notice that the knowl-
edge, be it generic or particular, is almost never overt. It may even be
anatomically implemented (Riedl 1988; Vollmer 1990). It always derives
from uncontradicted experience though, whether phylogenetically or
ontogenetically.
Because generic ecological optics (Gibson 1950) coincides with a subset
of physics, it is the same for all observers (species) and can be formally
understood via an analysis of the pertinent physics. Much of our present
knowledge derives from applied physics and computer vision. Whether
a given observer or species actually exploits such knowledge is of course
up to behavioral verification. Empirically we find that many generic facts
of ecological optics are reflected in the behavior of very different species,
though the precise implementations may differ widely. After all, flies are
constructed differently from man, though both orient and navigate partly
244 Jan J. Koenderink and Andrea J. van Doorn

via “optical flow.” This indicates, by the way, that a mindless reduction-
ism is not going to get one very far in understanding such feats of living
organisms.

Types of “Pictures”
There exists a bewildering variety of artifacts that are commonly referred
to as “pictures.” In this chapter, we reserve the term “images” for percep-
tions and use “pictures” for such things as photographs, paintings, and
drawings. Although pictures are simply physical objects, they are pictures
to an observer. That is to say, an observer has to take them as a picture. A
footprint in the sand is a picture to human observers, whereas a photo-
graph is merely a piece of paper to your dog. Images are in the mind,
whereas pictures may be handled. Here we are mainly interested in pic-
tures “of” something, either of actually existing things, such as a horse
or a cow, or imaginary things that still fit into some observer’s realm of
expertise in ecological optics. Think of unicorns or dragons, for instance.
For our purposes it is of minor interest whether a picture has been pro-
duced by some mechanical means such as footprints on the beach, photo-
graphs, or computer graphics, or whether it came about through an act of
artistic rendering. Indeed, it is impossible to decide on the basis of the pic-
ture as object alone. What will be important is the particular bouquet of
cues that the picture offers the observer. For instance, a silhouette, a car-
toon (line) drawing, a painting in flat tones, a pointillist (realistic) paint-
ing, and a straight photograph offer quite different bouquets of cues (see
figures 12.1 and 12.2). Remember that cues are bits of information and
exist only in relation to certain observers.
Because cues can be, and often are, qualitatively different chunks of
information, different pictures offer not simply different amounts of infor-
mation but qualitatively different informational contents. Thus different
renderings of a single “fiducial object” may well specify distinct pictorial
objects to an observer (see figure 12.2). The fiducial object may be a real
(or physical) object in the case of photographs, or a concept in the head
of an artist (a unicorn, say) that may materialize as a painting, drawing,
and so forth.
In many cases it may not be very useful to speak of a fiducial object at
all, because artists often aim at a concrete result, such as a painting, rather
than (or in addition to) the expression of an idea. Whereas many ideas can
be drawn, painted, or described in words, this need not be universally
the case. For instance, painted unicorns cannot be drawn, for the simple
245 Pictorial Space

Figure 12.1
Four pictures with identical geometry but different rendering.
246 Jan J. Koenderink and Andrea J. van Doorn

Subject CC
Silhouette Cartoon

Illumination Illumination
from lower right from upper left
247 Pictorial Space

reason that drawings are not paintings. In such cases, like in the case of a
face seen in a cloud, the picture is all there is.

Modes of Seeing
Pictorial spaces are mental entities and depend upon the observer just
as well as upon the picture. Without observers there are no images and
pictures remain forever “planar objects covered with pigments in cer-
tain simultaneous orders”; to echo Maurice Denis: “Remember that a
picture—before being a battle horse, a nude woman, or some anecdote—
is essentially a plane surface covered with colors assembled in a certain
order” (Denis 1890/1976, p. 380). There are many different modes of see-
ing: a face may be seen in a cloud, a person may be seen frontally (facing
you) in a painting viewed obliquely, a photograph may be seen as a flat
object and you may see a pictorial space in it simultaneously, and so forth.
Of particular interest are cases in which the observer assumes a certain
viewing mode with the express purpose to influence the structure of picto-
rial space. Such viewing methods are often practiced by artists (Leonardo
da Vinci 1989) to either flatten optical space for purposes of drawing, or
deepen pictorial relief for purposes of monitoring the effectiveness of ren-
dering. Optical space is flattened through such devices as monocular view-
ing, “screwing up the eyes,” viewing through a wireframe, using a Claude
glass (tinted mirror) or camera obscura, and so forth (Dubery and Willats
1972). (See figure 12.3.) Pictorial space is deepened through monocular
viewing, deemphasis of the frame, a frontoparallel vantage point, and so
forth.
The case of deepening the relief in pictorial space is interesting, because
the structure relating to the pictorial relief that is sampled by the eye(s) is
obviously the same whether you view the picture monocularly or binocu-
larly. Yet the “information” is different because the structure relating to
the actual scene—you holding the photograph and looking at it—is dif-
ferent. With two eyes open, there are strong indications that you are look-
ing at a flat object. Apparently this cue “conflicts” with the pictorial cues
proper. More on that later. This conflict is an indication that a clean

Figure 12.2
Pictorial relief for the four pictures shown in figure 12.1 with identical geometry
but different rendering. The lines are loci of constant depth. The pictures represent
different pictorial objects to the observer CC. In the case of the silhouette you may
guess the object is a torso, but it is ambiguous whether you see the front or back
of it. This observer apparently split the difference.
248 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.3
(Left) “framing” the scene in front of you and looking at and through the frame
with one eye is a time-honored method to “flatten” a scene. Optical space starts to
look as a picture. (Right) cutting out context is an excellent way to remove depth
and relief in order to be able to focus on textural properties and color.

segregation between “pictorial space,” that is, the space in which the pic-
torial content unfolds, and “physical space,” that is, the space that con-
tains the picture, is not necessarily made by a given observer.

The Issue of “Veridicality”


The issue of “veridicality” need not come up when you see a dragon in pic-
torial space or a face in a cloud, that is, when you are convinced you are
experiencing an “illusion,” but it seems natural enough in the context of
optical space or the pictorial space related to a straight photograph. Yet
the issue is a particularly vexing one, even in those cases. It is also an
important issue and is generally ignored or misunderstood.
The problem is that a photograph by no means specifies a scene,
although most people seem confident that it does. There exists a many-to-
one relation between scenes and photographs (see figure 12.4). In fact, a
given photograph could have been due to infinities of different scenes. One
particular example is the actual scene in front of the photographer at
the time of exposure. Let’s agree to call it the “fiducial scene.” Another is
the photograph itself in front of a camera, at the right distance and so
249 Pictorial Space

space of scenes space of percepts


(in the world) (in the mind)

A P

Expertise
in
ecological optics

I
space of images

Figure 12.4
The actual scene A gives rise to an actual image I, which again gives rise to a
veridical perception P. A “nonveridical” perception would indicate a scene that
would lead to a different picture. All scenes (like M) that lead to the picture I are
“metameric” to the actual scene and would be equally veridical perceptions. All
such metameric scenes are related by a group of ambiguity transformations.

forth. But many really different possibilities exist. It is perhaps sufficient


to remember that Hollywood movie producers rarely go to the trouble
to build castles or Martian landscapes. These are “faked,” though we all
know that the movie scenes can look only too real to us. It is easy enough
to think of the weirdest examples. For instance, consider a stellar constel-
lation that happens—by an incredible cosmic coincidence—to excite your
retina in the same way as when you look at Leonardo’s Gioconda at the
Louvre from a particular vantage point. The “pixels” are stars at essen-
tially random distances (many light years away). Though perhaps some-
what unlikely, this is clearly possible, and it vividly illustrates that things
need not at all be what they seem. George Berkeley (1709/1975) was quite
right that vision is impossible in principle! Helmholtz (1896) was right that
we see what is (for us) the most likely explanation of our present retinal
activity. The fact that we can write this shows that the odds are perhaps
slightly in our favor. Although vision is indeed impossible in principle, it
generally works fine for us.
Clearly observers “have to bring something” with them in order to
even stand a chance to look at a picture and see a pictorial space. This
250 Jan J. Koenderink and Andrea J. van Doorn

something is an intuitive understanding of the causal nexus of laws of


ecological optics (in a way “vision” is a process that embodies this under-
standing) and a generic knowledge of frequently encountered objects and
scenes. In the top left of figure 12.2 we see that the observer “fills in the
blanks” and because there is nothing in the picture except a global refer-
ence, the filled-in structure has to come from the observer’s memory.
Notice that any additional information is used to override the idiosyn-
cratic content, though. When we correlate the psychophysical response
reliefs of different observers we find that they correlate low on the silhou-
ette rendering, higher on the cartoon rendering, and very high on the
fully shaded renderings. This shows that the idiosyncratic content dimin-
ishes when information is added and that it is overridden by objectively
specified structure. The being overridden is the same for all observers
because it is due to structure that is in the picture and not in the head so
to speak.
When the reader pages forward, it will appear that we often use human
torsos as objects in our experiments. We do this because these are suitably
“generic smooth” objects as opposed to the cuboids, cylinders, and ellip-
soids that seem to dominate the literature. The latter are ill suited for the
study of pictorial relief because they are highly nongeneric and degener-
ate. Many algorithms from ecological optics will not work or malfunction
on such degenerate objects, whereas they work fine on generic ones. (For
instance, the contours of ellipsoids are planar curves, so are the isophotes
of shading. This is highly nongeneric.) When we present our work, there
is invariably someone who will object that “we know what humans look
like.” This is true to some extent (of course we also know what cylinders
and so forth look like). However, this “knowledge” seems to be mostly
limited to the ability to name parts (limbs, major bones, and muscles). As
anyone who ever attended a figure-drawing class knows, beginning stu-
dents have surprisingly scant understanding of the human form. Anyhow,
when we vary cues (for instance, shading) slightly, we notice changes in
the pictorial relief. Apparently observers easily let pictorial cues override
their superior knowledge of the human body. A study with computer-
rendered random shapes (potato-like blobs with various bulges and cavi-
ties) confirms that the use of these torsos should give no cause for concern.
Because the fiducial scene is by no means singled out among the infinite
set of scenes that could have led to the photograph, the issue of “veridi-
cality” is really a nonissue. This is equally true for optical space, because
there is hardly a difference between looking at pictures or at real scenes at
251 Pictorial Space

the level of the radiance. Veridicality in perception becomes an interesting


issue in the setting of perception and action—perhaps potential action—
where the issue becomes one of the success or failure of optically guided,
actual or potential, behavior. We strongly believe that the topic of veridi-
cality can be safely ignored when the intrinsic structure of pictorial space
is discussed. However, the reader be warned that we have been severely
criticized for this offbeat standpoint.
Recent progress in the analysis of optical information (“cues”) has
revealed the structure of the set of scenes that correspond to certain cues
(Belhumeur, Kriegman, and Yuille 1999; Faugeras 1993; Koenderink and
van Doorn 1997). Such analyses are only available for a small number
of cues and rather strong assumptions to constrain the problem and make
it tractable. We find that the “solutions” are mutually related by certain
“ambiguity transformations” that change the scene but leave the pho-
tograph (the bouquet of cues) invariant. Such transformed scenes are
obviously equally “valid” interpretations of the photograph. These ambi-
guities are not (indeed, by their very definition) specified by the photo-
graph, but the observer has perfect freedom to select one and thus arrive
at a unique interpretation of the photograph. Such selection is by its
nature fully idiosyncratic. We will refer to it (as in Gombrich 1959) as
the “beholder’s share.”

12.2 METHODS TO OPERATIONALIZE PICTORIAL RELIEF

The first problem we have to face in the study of “pictorial relief” is how
to “measure it.” The way we phrase it here perhaps sounds reasonable but
actually puts you on the wrong foot. The problem is that in order to meas-
ure something that “something” has to exist independently of the act of
measuring. If the something is operationally defined (by the measurement)
you don’t measure it, but rather the measurement creates “it.”
It is a priori evident that methods of measuring physical quantities will
not be applicable to the case of pictorial space, because pictorial space is
a mental entity. Likewise, physiological methods (electrophysiology, func-
tional brain imaging, etc.) are by their nature unable to address the prob-
lem. Pictorial relief is a mental entity and thus can only be probed via
mental operations. An example would be to have an observer describe the
pictorial space verbally or in terms of an artistic statement. Although such
methods certainly have their advantages, they are hardly suited to arrive
at a body of quantitative, parametric data.
252 Jan J. Koenderink and Andrea J. van Doorn

At the next level we may ask the observer for magnitude estimates
of “depth” or “surface attitude” (slant and tilt) and so forth of pictorial
points or surface elements. Such (verbal) estimates are judged to be
extremely difficult by typical observers and, what is worse, they yield very
variable responses (Mingolla and Todd 1986). Much easier are judgments
of relative magnitudes, such as “nearer” or “more distant” (Koenderink,
van Doorn, and Kappers 1996). Here we may establish fairly well-defined
discrimination thresholds under various, parametrically varied, conditions.
Such methods helped to make up most of the classical corpus of scientific
data on the topic, and they remain important to this day.
A problem with these classical methods that is hardly recognized in the
literature is that (discrimination) thresholds don’t yield any handle on
what the observers actually experience. Another problem is that it is pro-
hibitive to collect the really rich data sets we desire, say the equivalent of
an elaborate description, such as a fairly fine triangulation, of pictorial
relief. These data are indeed desirable because they would enable us to
perform calculations of a differential geometric nature and thus relate
apparently disparate data sets. For instance, you might want to predict
curvature data from attitude data through differentiation. But the opera-
tion of differentiation requires a fine-grained description. There is pre-
ciously little one can do with sparse data. In figures 12.5 and 12.6 we show
an example of what we mean by a “rich data set.” Here the response is
essentially a surface.
What is needed are methods that go beyond thresholds, yet avoid the
problems that bedevil absolute magnitude estimates, and that are fast,
that is, enable us to collect hundreds or thousands of data points within a
reasonable time span, such as an hour. Indeed, a picture may be roughly
characterized through at least 10 by 10 (100) to a 100 by a 100 (10,000)
samples, and a single experimental session cannot last for more than
roughly an hour (that is less than 10,000 seconds). Thus we need methods
that allow us to collect a data point at least every few seconds for periods
up to not much more than an hour. This evidently eliminates both mag-
nitude judgments, where observers say, “hmm, . . . , well, . . .” for a minute
per data point, and threshold measurements, which require a full psycho-
metric curve—at least a 100 yes/no judgments—per data point. It cannot
be stressed enough that powerful methods to address pictorial relief have
to be based on methods that yield rich data, essentially “fields of esti-
mates.” For instance, a dense field of depth estimates can be considered
a “surface,” that is, a true geometrical entity, whereas a mere handful of
253 Pictorial Space

Figure 12.5
(Left) a stimulus with two superimposed gauge figures. Notice that the white one
“fits,” and the black one doesn’t. (Right) Result of one-half hour’s work: a dense
field of gauge figure settings.

depth values is nothing but a (geometrically) meaningless bunch of num-


bers. But—mirabile dictu—the literature doesn’t acknowledge this basic
fact.
Suitable methods are methods of reproduction, methods of adjustment,
and methods of fit. In a “method of adjustment” the observer is asked to
adjust some pointer such as to indicate a direction, curvature, and so on,
of pictorial relief (see figure 12.7, left). The pointer may be superimposed
on the picture at the location of the sample. For instance, an arrow at a
certain location may be rotated so as to indicate the direction of the depth
gradient of pictorial relief at the position in pictorial space corresponding
to the location of the arrow in the picture.
In a “method of reproduction” the observer has to “reproduce” some
pictorial entity in terms of another entity, perhaps in a different pictorial
space. For example, the observer may be asked to reproduce a depth dif-
ference as a stretch in the frontoparallel plane or to reproduce a section
of a pictorial surface with a plane that extends purely in depth in the
frontoparallel plane (Koenderink et al. 2000b). (See figure 12.7, center.)
254 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.6
The result from the attitude sampling can be integrated to a depth field (left) or a
“pictorial relief” (surface in three-dimensional space) (right).

Notice that the slider bar might as well have been plotted outside the
picture frame proper.
In a “method of fit” an observer is asked to change the appearance of a
“gauge figure” in such a way that it “fits” the pictorial relief (see figure
12.7, right).
The method of fit is especially interesting because the judgment of fit
applies to pictorial space proper and to nothing external to it. This is not
true for the other methods, which introduce alien elements (markers,
pointers, etc.) that don’t properly belong to pictorial space but are graphi-
cal elements superimposed upon the picture that may (like the punc-
tate markers or the arrow) or may not (like the slider bar) relate to the
graphics sustaining pictorial space.
An instance of the method of fit that we have exploited frequently
(the method of fit is much more generally applicable, though) is the meas-
urement of surface attitude (Koenderink, van Doorn, and Kappers 1992).
(See figures 12.7, right, and figure 12.8.) We superimpose a gauge figure
that is under control of the observer, who is supposed to adjust it in such
255 Pictorial Space

Figure 12.7
(Left) example of a method of adjustment. Notice that the arrow exists in the pic-
ture plane, not in pictorial space. (Center) example of a method of reproduction.
Although the punctate markers seem to lie on the pictorial surface, the slider bar
is in the picture plane and seems fully unrelated to pictorial space. (Right) exam-
ple of a method of fit. The “gauge figure” belongs to pictorial space and relates to
pictorial objects.

a way that it “clings to the pictorial surface.” From the setting we derive
numerical values of slant and tilt, but at no time the observer is required
to estimate or even distinguish between slant and tilt. The task is simply
to establish a fit in pictorial space. Observers find this an easy (almost
trivial) task, whereas they are hard put to estimate slant and tilt values
when asked to do so.
We belabor this point because it is so often misconstrued. It is not that
the observer has to estimate the attitude of the pictorial surface and of the
gauge figure separately and somehow make them become equal. The task
is simply to produce a fit, and the slant and tilt never enter into this. The
fit can be judged cheerfully within a small fraction of a second, whereas
observers take up to a minute to estimate slant and tilt explicitly—and bit-
terly complain about that task. What is important here is that the method
of fit requires only judgments that are intrinsic to pictorial space. This
important fact is generally misunderstood. For instance, people have
256 Jan J. Koenderink and Andrea J. van Doorn

Ti
lt
Slant

Figure 12.8
Gauge figures used to probe spatial attitude of pictorial surface elements.
Parameters are the slant and tilt angles. The array shows a “natural” planar inter-
face: a polar diagram with slant as radius and tilt as orientation.

studied the judgment of slant and tilt of a gauge figure in isolation and
used the result to “correct” the results of the fit method. It is entirely
unclear that such a procedure is at all reasonable unless one is ready
to take very strong and—to our mind—unlikely prior assumptions for
granted. It seems much more reasonable to take the results of the method
of fit as an “operational definition” of pictorial relief and to take the
work on gauge figures in isolation as (perhaps interesting) essentially un-
related data. In order to relate such initially unrelated data, one stands in
need of some theory. The theory itself is subject to empirical verification
of course.

Local and Global


An important distinction to be made is that between judgments of prop-
erties that exist at a single location of pictorial space, we call these “local”
properties, and judgments that require the comparison of properties that
simultaneously exist at two or more disparate locations, we will refer to
these as “global” properties. The difference is important for the practical
reason that observers tend to have little problems with local tasks but find
global tasks extremely difficult to perform. It is also important because of
conceptual reasons. It is not at all clear that freedom to compare proper-
ties at different locations in pictorial space should exist at all. This is up
257 Pictorial Space

P P P
T Q

Figure 12.9
The relief in the neighborhood of a point P in the zeroth-order (leftmost, only
location), first-order (center, tangent plane T added), and second-order (right-
most, “osculating quadric” Q added) approximation.

to empirical verification. After all “pictorial space” is virgin territory and


it is quite unknown whether objects can be moved about and brought
into close vicinity of each other for the sake of easy comparison, with the
mad abandon they can in Euclidian space. We certainly should avoid far-
reaching prior assumptions here. Otherwise the conclusions aren’t worth
more than the assumptions.

Orders of Differentiation
When “pictorial relief” is a more or less smooth surface, like the surface
of a polished marble sculpture, then it should be possible to speak sen-
sibly of various “orders of approximation” to the surface (Koenderink
1990). (See figure 12.9.) In the “zeroth order” we merely indicate the loca-
tion of a point on the surface. Thus we know where the surface is, but we
don’t know its attitude, let alone its curvature, and so forth. In the “first-
order” approximation we also know the attitude of the tangent plane at
the point. Then we know both the location and the orientation of the sur-
face, but not its curvature, and so forth. In the “second-order” approxi-
mation we know in addition the “osculating quadric” of the surface. Then
we know the location, attitude, and (local) shape of the surface. The local
shape is defined via the directions of principal curvature and these princi-
pal curvatures themselves, and so forth, for the still higher orders.
When we indicate a point in the picture (two degrees of freedom, e.g.,
Cartesian coordinates), the zeroth order gives another degree of freedom
(the depth, a third Cartesian coordinate), the first order adds two degrees
of freedom (e.g., the slant and tilt angles of the surface), and the second
order adds another three degrees of freedom (the direction of the maxi-
mum curvature and the two principal curvatures, e.g.).
258 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.10
“Integratability constraints.” On the left, the “deck of cards” example. Second
from the left, a case where a field of tangent planes can easily be integrated to a
closed ribbon. For the case second from the right, this doesn’t work (the planar
elements at the endpoints are at different depths). The case on the right is differ-
ent: the planar facets at the endpoints have different spatial attitudes. In order to
obtain a “fit,” the field of planar elements has to be quite special.

Of course the observer might also be sensitive to a higher-order (say,


curvature) and insensitive to lower order(s) (for instance, depth and slant).
This is up to empirical verification. When such would turn out to be
the case, there should be little reason for surprise. After all, the human
observer is exquisitely sensitive to luminance contrast but rather insen-
sitive to the absolute luminance level. Cases like these are frequent in
human perception.
Reference to the order of approximation is very practical, because most
methods yield estimates of properties of only a certain order. Estimates
of depth sample only the zeroth order, the method of fit described earlier
samples the first order, and so forth. Notice that the order of approxima-
tion applies only to purely local methods.
By differentiation of a field of local estimates, you obtain a prediction
for the next higher order (thus depth estimates allow a prediction of sur-
face slant and tilt). By integration of a field of estimates, you obtain the
next lower order up to indeterminate additive constants. Thus surface atti-
tude estimates allow the prediction of a depth field up to some unknown
depth shift. In the latter case there is a little, but important, twist in that
it is not necessarily possible to perform the integration at all, unless the
initial estimates satisfy some constraint (Spivak 1975). For instance, the
playing cards in a deck are “local attitude estimates” (see figure 12.10, left)
but clearly don’t allow of an integral surface. In this case it means that you
gain a check on “surface consistency” (see figure 12.10).
Differentiation takes you from global to local, whereas integration
takes you from local to global. Thus there exist a great many possibilities
259 Pictorial Space

of comparison of apparently qualitatively different results. The important


condition is that you need rich enough data, that is to say, “fields” of esti-
mates, rather than mere point samples. This is the main reason the classi-
cal results have remained essentially barren. There is simply no way to
compare results of different experiments in interesting ways. Of course the
literature doesn’t acknowledge this fundamental fact, because it is not
concerned with rich data sets in the first place.

Cognition versus Perception


“Perception”—visual perception, that is—is best defined as “optically
guided potential behavior.” Notice that a well-known definition in biol-
ogy “optically guided behavior” will not do for us, because it admits pure
reflex action as “vision” and has no place for consciousness, which is why
our definition was framed as it is of course. Because pictorial space is a
mental entity, we cannot talk about it except in the context of conscious-
ness. Our definition manages to catch that, yet remains squarely in the
realm of objective science.
Potential behavior can be probed through variation of the observer’s
goals. When the behavior correlates with the optical structure over arbi-
trary goal variations, we may say that the observer behaves in a certain
way because he or she has certain goals and because he or she knew cer-
tain facts due to optical interaction (“saw something”). Notice that this is
simply the commonsense definition. When you hear the public address
system at an airport make some announcement and see people do various
things that may differ according to their goals in response to the sound,
you know they both heard and understood the announcement and did not
react reflexively, as they might have in response to the sound of a sudden
explosion. Reflex actions are stereotypical, whereas actualized potential
actions can be virtually anything, depending on the present goals. It might
seem a drawback of the definition that it applies only to observers for
whom you can influence the goals, for example, by promising payment
in return for their efforts. Thus you can’t probe visual perception of
Martians or mollusks, although these may well display optically guided
behavior. Indeed, psychophysics, as opposed to physiology, implies a cer-
tain degree of “empathy.”
Though observers may show indications of visual perception, they need
not necessarily be able to verbalize their percepts, nor need they be able to
use the percepts in overt reasoning. For example, cats and dogs clearly
perceive yet are not capable of such cognitive acts. In fact, the result of
260 Jan J. Koenderink and Andrea J. van Doorn

overt reasoning is not a perception in the proper sense. This is an impor-


tant point because some methods require overt reasoning or verbaliza-
tion (e.g., “call out slant and tilt in degrees for yonder surface patch”)
whereas others don’t (e.g., “make that gauge figure fit that surface”).
Moreover, some observers may be rather immediately “optically guided,”
whereas others seem to reason out what to do. For instance, in the gauge
figure fitting task they might actually make overt estimates in degrees, rea-
son out the shape of the gauge figure for those estimates, then change the
gauge figure to that form.
We find (purely empirically) that observers typically fall into two dis-
tinct classes. Observers from one class take a second or so for a gauge
figure setting, the others at least 10 to a 100 times as long. Practice doesn’t
change this, the “fast” ones don’t need any practice at all, and the “slow”
ones never “learn” anyway. This observation strongly suggests that such
observers function quite differently. Moreover, some implementations of
certain tasks almost force observers to switch from perceptual to cognitive
strategies. For instance, when you interface the attitude gauge figure in
such a way that the observer has to set slant/tilt in Cartesian coordi-
nates instead of the more natural polar way (Pollick et al. 1992), the task
becomes an almost impossible one and requires much careful thinking.
Ever try to write your name with an XY-movement device? Though such
details may seem irrelevant, indeed often go unmentioned in publications,
they are really crucial.
In our work we carefully optimize the interfaces and we select observers
(simply by ignoring the—very small—group that takes at least an order
of magnitude longer at the task than most others will) to ensure that we
address perceptual, rather than cognitive, issues. As a rule of thumb a
setting shouldn’t take more than a few seconds at most and should feel
“natural” so that the observer almost effortlessly keeps at the task for a
half hour at a stretch. In such cases as when settings take longer than a few
seconds at most and observers complain that something is seriously awry,
we would hesitate to accept the results as very informative.
It makes sense to check for such things in the method sections of papers,
though the issue is rarely addressed. Yet any reasoning on the part of
the observer may introduce spurious elements (see later discussion). This
consideration, although rarely mentioned, is far more important than the
number of observers, whether they are naive, and so on, factors based on
which reviewers of research papers conventionally rate the scientific merit
of manuscripts.
261 Pictorial Space

The Meaning of a “Measurement”


When you have to measure the temperature in a bucket of water you are
for once in good shape. For physics has developed a theory of many types
of potential thermometers and—even more important—you may rest
assured that “the temperature” exists even if you don’t measure it and
that various methods to measure will lead to the same result for that very
reason. Things can work out quite differently in other fields. For instance,
the measurement of cosmological distances is difficult because the con-
cept of “distance” may be different on scales we’re not accustomed to and
because the various methods sometimes lead to different results. If so,
which one is right? Is any of them right? Does “being right” have a mean-
ing at all?
The measurement of properties of pictorial space are in yet another cat-
egory. It is by no means a priori clear that “surface attitude,” say, exists
before the act of measurement. The act of measurement creates a novel
state of affairs for the observer that yields a certain result, for instance, a
number. Thus different methods to measure it need not refer to the same
property, at least not in principle; this is up to empirical verification. The
numbers you obtain should be labeled with the method by which they
were obtained, for they are only operationally defined.
Failure to grasp this basic point has led to numerous misunderstandings
in the literature where things are typically labeled with familiar terms
(“depth,” “slant,” etc.) without any further indexing by method. But there
may exist numerous “depths” and there is no a priori reason they should
correspond. To think that they “should” is to assume that pictorial space
exists in the same way that a bucket of water does. That is clearly prepos-
terous, at least to us. Many experts in the field think differently though,
the reader be warned.
Notice that this argument indicates that we should not look for any
single “best” method to measure pictorial relief. For best has no meaning
here. Instead, it seems most appropriate to work with as extensive a bat-
tery of methods as possible and to look upon the comparison of different
methods to measure “the same” entity as a valuable research tool by itself.
Rather similar problems arise when measurements of different proper-
ties are compared. For instance, although you can indeed obtain surface
attitude by differentiation of a depth field in Euclidian space, the same
operation may turn out to be baseless in pictorial space. Again, these are
empirical issues. Pictorial space is an entity to be explored, and we should
be wary of attributing properties to it (most probably similar to those of
262 Jan J. Koenderink and Andrea J. van Doorn

our trusty Euclidian space) without explicit empirical evidence. Again,


ignoring this basic point has led to numerous unfortunate misunder-
standings in the literature. This is also one reason observers in the
“cognitive mode” may yield results that differ from those from observers
in the “perceptual” mode. The former observers typically apply (perhaps
unknowingly) knowledge of Euclidian relations to change the task into
“equivalent” tasks that they find easier to perform. For instance, they may
assume that two angles of an isosceles triangle have to be equal, although
there is no a priori reason this should be the case in the visual field.
That we seem to be consciously aware of globally consistent pictorial
surfaces when we look into a straight photograph of a piece of sculpture
is a feeling that need not necessarily be corroborated empirically. It may
well be the case that various, mutually incompatible, representations are
simultaneously present in our subconscious; that our awareness merely
glosses over this, not resolving ambiguity when the task doesn’t call for
it; and that our awareness simply selects an appropriate representation
for the task at hand. Then the inconsistencies would not necessarily be
noticed, and results obtained by different methods may well turn out to be
mutually inconsistent. If such possibilities exist, and indeed, neurophysio-
logical findings seem to indicate that this is most likely, then the result of
a measurement need not exist in the mind before the very act of measure-
ment. We are at far remove indeed from the measurement of the tem-
perature of a bucket of water.

12.3 PRELIMINARY EMPIRICAL FINDINGS

A simple method to operationalize pictorial relief is by way of surface


attitude measurements (see figures 12.5 and 12.6). Observers can easily
produce 1,000 samples per hour with the gauge figure fitting method. In
practice we perform measurements at the barycenters of a regular triangu-
lation. The triangulation remains hidden to the observer. Measurements
are done in random order, one at a time. This produces the attitude or
depth gradient of the faces of the triangulation. We then assign depth
values to the vertices in such a triangulation in order to “explain” the
depth gradients of the faces in the best possible way. These depth values
are only determined up to an arbitrary common constant. The residual
difference between the depth gradients explained by the depths at the ver-
tices and the actual measurements may be due to either of two factors.
One is the intrinsic uncertainty in the measurements, which can easily be
263 Pictorial Space

AD AK

Figure 12.11
On the left, the stimulus; on the right, two pictorial reliefs from different observers.
Depth runs to the right. Notice that the reliefs are different and differ mainly by a
depth stretch.

estimated from the scatter in repeated trials. The other is a possible


“inconsistency.” The empirical depth gradients may actually turn out
to be incompatible with any surface. We have found that the residuals
can be explained from the scatter in repeated trials. Thus (most remark-
ably) the local attitude samples are invariably compatible with a sur-
face. This “pictorial relief” thus explains the results from local attitude
samples.
We may consider the pictorial relief to be the “psychophysical response”
to the stimulus, that is, the picture. Two pictorial reliefs may be compared
in many different ways. A way that is often useful is to simply look at a
scatter plot of all the depth values at corresponding points in the picture.
When the correlation is high, the slope is a relevant entity; it specifies
a depth magnification between the two reliefs. We find that when we test
the same observer repeatedly over a long period, the coefficients of deter-
mination (“R-squares”) are in the 99% ballpark and slopes vary unpre-
dictably by factors of up to 15%. When different observers are compared,
the coefficients of determination are also quite high (R-squares typically
over 95%), but the slopes may differ by a factor of two or more. What
this means is that observers typically generate the same pictorial relief
modulo an idiosyncratic depth stretch (figure 12.11). With respect to the
notion of veridicality, we conclude that the responses of at least one (but
probably both) observers cannot be veridical in the naive sense.
264 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.12
On the left, the stimulus, and in the center, a pictorial relief with depth increasing
toward the right. On the right, a picture of the actual scene taken at right angles
with respect to the viewing direction for the stimulus. “Naive veridicality” would
imply identity of the outlines of the relief (center) and picture of the rotated object
(right). Clearly it does not pertain.

As expected, the pictorial reliefs we find are superficially similar to the


actual object at the time of exposure of the photograph but (often very)
significantly different in detail (figure 12.12). Notice that we cannot draw
any conclusions as to veridicality from this, because there is no easy way
to find out whether the pictorial reliefs generated by the observers indicate
objects that might have yielded the photograph. In a trivial way they
always might; that is, when you allow the object to be painted ad libitum,
then a trompe l’oeil painting would fit the bill. Thus the problem has to be
suitably constrained in order to have it make sense at all. We consider the
issue of veridicality irrelevant in this context.

Viewing Modes
When a method such as the one discussed yields a pictorial relief as psy-
chophysical response to a photograph (the stimulus), we may study the
stimulus-response relation over a large number of parametric variations
of both the stimulus and the viewing conditions. In this section we con-
sider the variation of viewing conditions.
265 Pictorial Space

Figure 12.13
Stimulus and three pictorial reliefs, from left to right, with binocular, monocular,
and synoptical (see figure 12.14) viewing. In these pictures the depth dimension
is vertically upward. Notice the differences in total depth range and qualitative
similarity of shape.

When only the viewing conditions are changed, the photograph is invari-
ant and so are all cues that relate to pictorial space. It is only the cues
that relate to the photograph as an object in optical space that are varied.
In conventional language we may speak of a “cue conflict” situation.
Because the observers are always simultaneously aware of both the opti-
cal space that contains the photograph as a physical object and the pic-
torial space that exists as a completely disparate entity, cue conflict is
perhaps an unfortunate term, as it suggests that different cues are in
conflict with respect to the resolution of a single entity. However that
may be, it is evident that a change of viewing conditions, such as viewing
monocularly versus binocularly, looking straight at the photograph versus
viewing it at an oblique angle, and so forth, affects the cues that relate to
the photograph as a flat object in optical space. They do not affect the cues
that relate to pictorial space.
We find that such changes in viewing conditions have a strong influence
on pictorial relief (see figures 12.13 and 12.14). As perhaps expected, the
relief remains qualitatively the same—after all, the pictorial cues are iden-
tical. What happens is that pictorial space suffers a “depth stretch,” whose
amount is immediately correlated with the viewing condition. We find that
the depth of pictorial relief is expanded by factors up to two when the
observer changes from binocular to monocular vision by closing one eye,
assuming normal binocular stereopsis. Similarly, we find that depth of
266 Jan J. Koenderink and Andrea J. van Doorn

binocular depth
200 synoptical depth

100

-100 monocular depth

-100 0 100 200


Figure 12.14
(Left) the principle of the “Synopter.” Left and right eyes are optically super-
imposed into a single “cyclopean eye.” (Right) Scatter plot of depths at corre-
sponding points of the reliefs shown in figure 12.13. Notice the nice, clean linear
correlation. The differences are a mere depth dilation or contraction.

pictorial relief increases by factors up to two depending upon the obliquity


when the observer moves from an oblique vantage point to a normal one.
When we allow the observer to look with both eyes but introduce an
optical mechanism that removes the disparity by superimposing the cen-
ters of rotation of the eyes optically, depth of relief increases by factors up
to two over that obtained with monocular viewing. With such “synoptic
viewing,” we have found increases of up to a factor of five over binocular
viewing (Koenderink, van Doorn, and Kappers 1994). (See figures 12.13
and 12.14.) Indeed, the change in pictorial relief brought about by such a
“synopter” is a striking one and has been commented on before (Carl
Zeiss Jena 1907).

Cue Variation
When we fix viewing conditions, for example, monocular viewing, and
vary the photograph we expect qualitative changes of the pictorial relief.
If such changes do not occur, or are only minor, we may well use conven-
tional terminology and speak of “constancy.”
267 Pictorial Space

A particularly interesting class of experiments is that in which the geom-


etry of the scene is kept fixed. We use photographs of an object with fixed
pose and fixed camera position. Changes in the resulting photograph,
then, are brought about through changes in lighting, background, or sur-
face properties of the object. In such cases the changes in pictorial relief
can easily be correlated with changes in the scene.
In one particular study (figures 12.15 through 12.17) we change the
direction of the main light source in the scene. The resulting photographs
look greatly different, though geometrically identical. We find that the
pictorial reliefs are very similar despite the differences in appearance, thus
revealing a remarkable “shape constancy” with respect to large varia-
tions in shading. However, closer analysis reveals that—superimposed
over the constancy—there remain minor but clearly significant variations
in the relief that are correlated with the changes in illumination direction.
These residuals are qualitative changes of the relief, rather than mere vari-
ations of the depth of relief. An intuitive way to describe the nature of
these changes is to say that protrusions of the relief tend to point in the
direction of the light source.
As a general point of methodology, we find that such parametric varia-
tion of a single cue as one component of a full bouquet of cues, and sub-
sequent correlation of the response with the parametric variation, is a
much better way to study the importance of particular cues than the
traditional reductionist method of cue isolation. Many cues cannot be
studied properly in isolation at all, because they only contain useful infor-
mation in a setting that includes other cues. An instance is the shading cue
just referred to, which is essentially inoperative in the absence of contours
(Erens, Kappers, and Koenderink 1991, 1993). The nature of a specific
cue is most properly studied when the “set point,” to borrow a term from
systems engineering, is a natural one. This is an important consideration
because most of the literature on cues either uses the method of reduction,
that is, to bring cues in isolation, or cue conflict, that is, to “pit cues
against each other.” Neither method allows one to confront the observer
with pictures that would ever be considered anything close to “normal”
by naive people. Common sense suggests that it may be prudent to prefer
the more “natural” stimuli.

Local versus Global


The pictorial relief as operationalized via the gauge figure set method is an
example of a global geometrical entity, a surface in a three-dimensional
268 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.15
Four stimuli with identical geometry but different chiaroscuro. The illumination is
from top left in the top left figure; in the top right figure it is from the top right; in
the bottom left figure it is from bottom left; and the bottom right figure is illumi-
nated from the bottom right.
269 Pictorial Space

UL UR

LL LR

Figure 12.16
The pictorial reliefs obtained from the stimuli shown in figure 12.15 for a single
observer. The curves are loci of constant pictorial depth.
270 Jan J. Koenderink and Andrea J. van Doorn

UL UR

LL LR

Figure 12.17
The residuals for the pictorial reliefs of a single observer (shown in figure 12.16)
obtained with the stimuli shown in figure 12.15. Notice that the distribution is
systematic, rather than random. The pattern of deviations from the mean of the
four reliefs depends systematically upon the direction of illumination of the scene.
271 Pictorial Space

space, that has been obtained via the integration of many (a field of) local
samples of surface attitudes. The local settings are the immediate result of
the observer’s efforts; the integration is done only after all settings are
completed, via a mechanical process of numerical integration. Remember
that the settings are obtained in random order, quite independently of
each other. It is an empirical question whether the surface as a global
entity is available to the observer, in the sense of being a causal factor in
the successful completion of some task, or not. A priori this availability
is not necessary to the observer’s success, even though all the necessary
data to determine the global surface (the individual settings) are evidently
available to the observer in the sense that they are produced when the
observer is asked to perform the gauge figure fitting task.
In order to establish whether the observer has access to the global sur-
face as an integral entity, we need a task of a global nature. A possibility
that we have explored (Koenderink and van Doorn 1995) is to put two
punctate marks on the photograph and let the observer decide which of
the marks appears nearer in depth. The marks are on the photograph, but
they are seen in pictorial space and they appear to be located on the picto-
rial surface. They don’t appear to hover in front of the pictorial surface,
for instance. Thus the task is accepted as a natural one by all observers.
When we fix one mark and vary the location of the other mark, we can
explore the probability of the observer’s judging the mark “nearer” in a
two-alternative forced choice task (“nearer/farther”). When the observer
has access to the integral surface, we expect the result to be almost binary.
For the mark should be closer with 100% probability if the mark on the
global surface is nearer than the fiducial mark, and it should be closer with
0% probability in case it is farther away. Only on the curve on the surface
that is the locus of points all equal in depth to the fiducial mark do we
expect probabilities of intermediate values. The task should be a trivial
one. Here is how to do it: simply threshold the integral surface at the
depth of the fiducial mark.
But this is not at all what we find (see figures 12.18 and 12.19). We find
that the depth order of any two marks can be established with some
degree of certainty only in cases where the points can be joined by a path
that either goes into depth or toward the observer, but not when such
paths don’t exist and there is a hill or valley that separates the points
(see figure 12.19). It is as if the observers can only check whether the direc-
tion of the gradient at nearby points matches up. Thus the pictorial relief
as a global geometrical entity is not available to the observers, even
272 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.18
(Left) A stimulus. (Right) The pictorial relief as measured via gauge figure fits.

though a field of local data that is sufficient to determine the global entity
was obtained psychophysically and is thus in principle available to the
observers.
This is a very important and perhaps somewhat surprising finding.
There is much information (or structure?) in the brain that is not available
to our consciousness. What enters the brain need not enter the mind.

Task Variation
When the task is varied for the same picture and viewing conditions, we
have a case in which the “freedom” of the observer is constrained by the
picture, that is to say, the bouquet of cues. Though we may expect differ-
ent results from different tasks, all these results should be compatible with
the cues. We have found a particularly surprising example (Koenderink
et al. 2000b) of a pair of similar tasks that turn out to yield spectacularly
different results (see below). We use a method of reproduction that requires
the observer to produce a (global) “normal section” of a pictorial object.
A normal section is defined as the intersection of the pictorial object with
a plane that contains the visual direction. In the picture, such a plane can
be represented through a straight line. In the paradigm we simply draw
such a line over the picture and let the observer produce the shape of the
273 Pictorial Space

Figure 12.19
(Left) Prediction of the probability for any point to be closer than the fiducial
point (gray hexagonal dot) arrived at by thresholding the pictorial relief obtained
from a gauge figure fitting method. (Center) Prediction on the basis of depth gra-
dient following. (Right) The actual probabilities obtained empirically. Notice that
the result is perhaps closer to the latter than the former prediction, although the
observer can do the global task to some extent.

intersection in a drawing (another picture) that represents the plane of the


section. Thus we have a paradigm where the response to a stimulus (a pic-
ture) is another picture, one that contains a curved line (the curve of inter-
section). The whole procedure is suitably constrained in order to enforce
reproducible results.
The lines in the stimulus picture may have any orientation. The differ-
ence in the two tasks lies in the orientation of the drawing. Notice that the
picture plane can be represented by a straight line in the drawing. In one
variation of the task, this line is always parallel to the line that represents
the cut in the picture. Here we mean parallelism in the visual field, which
contains both the stimulus picture and the drawing in a single plane. In
another variation of the task, the line that represents the picture plane in
the drawing is always horizontal (see figure 12.20).
We find, perhaps surprisingly, that a straight scatter plot of picto-
rial depths at corresponding points obtained with these two tasks reveals
no significant correlation (see figure 12.21, left). Thus the pictorial reliefs
obtained with the two tasks are not related by a depth stretch. When we
274 Jan J. Koenderink and Andrea J. van Doorn

adjust adjust

A B
Figure 12.20
The method of reproduction of normal sections. The observers have to drag points
in an auxiliary (A or B) orthogonally to the line.

60 60
B B
40 40

20 20

0 0

-20 -20

-40 -40

-60 A -60 A*
-60 -40 -20 0 20 40 60 -60 -40 -20 0 20 40 60

Figure 12.21
(Left) A straight scatter plot of the depths at corresponding points obtained with
methods A and B. (Right) Result of a multiple regression including the picture
plane coordinates.

perform a multivariate regression on the pictorial depths and the Cartesian


picture plane coordinates, we find very significant correlations with coeffi-
cients of determination of over 95% (see figure 12.21, right). This shows
that the pictorial reliefs may be different but also that the difference is of
a special kind. There exists an affine transformation such that the reliefs
are actually identical modulo the affinity (see figure 12.22).
In another example (figures 12.23 and 12.24) we show that the affinity
may be far removed from the identity, in this example a shear of more
than 60°! Thus the frontoparallel plane may be slanted by tens of degrees
by such an affinity. This is a most remarkable finding. It means that
275 Pictorial Space

A B A*

Figure 12.22
(Left) Pictorial relief (depth in the horizontal direction) obtained with method A.
(Center) Pictorial relief for the same stimulus, obtained with method B. These
reliefs produce the scatter plot shown in figure 12.21, left. (Right) The pictorial
relief shown on the left after an affine transformation obtained via multiple regres-
sion including the picture plane coordinates (figure 12.21, right). Now the reliefs
obtained with the two methods are nearly equal. Notice that the transformation
looks much like a rotation (although it is a depth shear).

Figure 12.23
On the left the stimulus, a “turtle” by Constantin Brancusi; on the right, the
(rather coarse) triangulation used in the study.
276 Jan J. Koenderink and Andrea J. van Doorn

100 100

0
0

-100
-100

-100 0 100 -100 0 100


Figure 12.24
The pictorial reliefs (curves of constant pictorial depth, top row) for the picture
shown in figure 12.23, left, and a single observer. Here the results are due to an
apparently minor variation of the cross-section task. Notice that the frontoparal-
lel points are completely different. In the bottom row we show a straight scatter
plot of depth (left) and one with an affine correction. The affinity has the effect of
raising the correlation from not significant to highly significant indeed.

even such a (at first blush well-defined and simple) geometrical prop-
erty as “frontoparallelity” need not be well defined, and indeed may be
extremely volatile with respect to minor task variations, in pictorial space.
Such a result should be shocking when you have been led to believe that
pictorial spaces due to photographs roughly resemble (in an Euclidean
way) the physical scenes in front of the camera at the moment of exposure.
They don’t, at least not in any naive sense.
We have found such transformations even in the case of single task
results. In such cases we sometimes find low correlations between observ-
ers. Scatter plots often reveal high correlations in certain areas of the pic-
ture. Apparently different transformations apply to different parts. An
277 Pictorial Space

Figure 12.25
A comparison of pictorial reliefs for two observers (same task, same picture). The
precise nature of the task is not important here; we find similar results with any
task. In the scatter plot of depths (on the right) we detect discrete branches. On
further analysis these correspond to contiguous regions (“parts”) in the picture
(figure at the left).

example is shown in figure 12.25. This observation indicates that things do


not necessarily go on in a global, but rather in a piecewise, fashion, the
pieces corresponding to “natural parts.” We have reason to believe that
observers use different “mental viewpoints” (see later discussion) for the
different parts. This strategy is reminescent of certain drawings by Picasso
(Bouret 1950), where this artist combines various perspectives in a single
drawing.

The Beholder’s Share


The “beholder’s share” is often an important factor in pictorial relief. We
distinguish three essentially different classes of pictorial relief in which the
beholder’s share plays a decisive or at least quantitatively significant role:
1. In illusions of the “faces in the clouds” variety, the beholder’s share
is decisive. It may be called “creative imagination,” perhaps somewhat
checked by structures in the picture.
2. In cases of pictures that lack certain pieces of vital information, the
observer’s share must “fill in the blanks” (see figures 12.1 and 12.2). This
278 Jan J. Koenderink and Andrea J. van Doorn

is also creative imagination, though much more thoroughly constrained


by the picture. An example would be a silhouette of a human body.
Although the picture may be clear enough in the sense that you see a
human body in it, it remains essentially indeterminate whether the picture
depicts a ventral or a dorsal view of a figure. Observers typically “per-
ceive” the one or the other, though, not some kind of entity that is simul-
taneously a ventral and a dorsal view.
3. In straight photographs that would generally be considered “true” and
“complete” pictures, the beholder’s share has much less freedom. It is
still the case that a picture is compatible with infinitely many physical
scenes, though, and the beholder’s share has certainly the freedom to
roam through this huge space of equivalent interpretations of the picture
and to select one as the present percept.
In the third case we may draw on the formal analysis of optical cues in
order to predict the freedom left to the beholder’s share.

12.4 THE STRUCTURE OF “PICTORIAL SPACE”

Does pictorial space have a similar structure to the space we move in? The
latter will be taken as the conventional three-dimensional Euclidian space.
Consider Euclidean space first. Euclidean space is infinitely extended and
looks essentially the same anywhere. Any line can be extended indefi-
nitely in either of its two directions and admits of an aperiodic, linear
metric. In contradistinction, the angular metric of the Euclidean plane is
periodic; the angles are confined to a finite range. In Euclidian space you
can make a full turn and end up in the same orientation. The difference
between the length and angle metrics precludes full duality between points
and planes. Although three distinct points determine a unique plane, three
different planes need not specify a unique point, because two of them may
turn out to be parallel. Although there exist parallel lines and planes, you
have no parallel points in Euclidian space, because any two points can be
joined with a line. Thus Euclidian space is much less symmetric than for
instance projective space. The group of orientation-preserving isometries
(also known as congruences, or motions) of Euclidian space contains rota-
tions and translations. “Euclidean shapes” are invariants under congru-
ences. By definition, two objects have “the same shape” when they can be
brought into superposition through a suitably chosen motion.
It is a priori evident that pictorial space is not at all like Euclidian space.
For instance, the “visual rays” are different in character from lines in fron-
toparallel planes. The length metric cannot be the same for these lines.
279 Pictorial Space

You cannot make a full turn in pictorial space, for it will be forever impos-
sible to obtain a dorsal view of a person (as pictorial object) when the
picture is a photograph in a ventral view. Thus the angle metric must
be aperiodic, like the length metric is. When we define pictorial lines as
straight, one-dimensional entities that correspond to straight lines in the
picture, then there exist “parallel points,” namely, points that cannot be
connected by any straight line. Such points coincide in the picture and
thus fail to define a line in the picture. They lie on a single visual ray. But
notice that the visual rays fail to count as proper pictorial lines because
they do not correspond to any line in the picture.
Thus pictorial space is quite different from Euclidian space, and it
becomes of much interest to establish its geometry in formal terms. The
next few sections are devoted to this issue.

Hildebrand’s Contribution
Hildebrand has to be considered the first person to establish a formal
framework for pictorial space. He noticed that observers often fail to dis-
tinguish between flat relief and sculpture in the round. Thus he concluded
that arbitrary depth dilations and contractions have to be considered
either congruences or similarities of pictorial space.
This idea has immediate consequences for the formal analysis of
“shapes” and “features” of pictorial objects. These must be invariants
under arbitrary depth scalings. We have studied these invariants previ-
ously: they are the “depth flow” (Hildebrand’s term, it means essentially
the congruence of curves of fastest depth descent on pictorial surfaces), the
near and far points (which are points on the pictorial relief where the local
tangent planes are frontoparallel), the ridges (like “divides” of the flow)
and ruts (the “courses” of the flow) (Koenderink and van Doorn 1995,
1998). Such features as ridges and ruts are indeed important features that
have evidently been deployed by many sculptors. We can easily compute
them for the pictorial reliefs we obtain empirically (see figure 12.26).

The Group of Congruences


Although Hildebrand evidently identified an important group of con-
gruences or similarities, we don’t think it is complete, and hence it is
insufficient to fully define the structure of pictorial space. Indeed, although
we have found numerous examples of the Hilbebrand type of ambigui-
ties, we have found impressive evidence of a much larger group of such
congruences, namely, the affinities mentioned earlier. It seems somewhat
surprising to us that Hildebrand failed to mention these, because relief
280 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.26
The Hildebrand invariants of pictorial relief in a particular case. Shown are near
points (circles with an open center dot), far points (circles with a filled center
dot), and saddle points (circles with a cross), as well as ridges (dark lines) and ruts
(gray lines). Near, far, and saddle points that don’t appear on the boundary are
frontoparallel points.

sculpture involving such affinities is not particularly hard to find in West-


ern art (Wittkower 1977).
A general argument that strongly suggests that the special group of
affinities mentioned earlier, which includes Hildebrand’s depth scalings,
exhausts the pictorial congruences is the following. We use the fact that all
cues define shape only up to some group of ambiguities (see figure 12.27).
For convenience we assume orthographic projection throughout.
The key observation is that almost all cues allow the observer to distin-
guish planar from nonplanar surfaces. A typical example is “shape from
shading.” In the usual setting, typically involving illumination with a
single parallel, collimated beam illuminating Lambertian surfaces, planes
give rise to uniform patches in the picture (Horn and Brooks 1989). Any
281 Pictorial Space

Figure 12.27
An example of the ambiguities associated with “shape from shading.” The convex
and the saddle-shaped surface patch can be illuminated in such ways (the illumi-
nation direction is indicated by the arrows) that a picture (taken vertically down-
ward) turns out to be identical for both of them (the figure on the right).

Figure 12.28
The patches on the left and center are uniform. The shape from shading inference
is that they are planar, though their spatial attitudes are indeterminate. The patch
on the right reveals some nonuniformity; thus it cannot be planar.

nonplanarity induces nonuniformity, such as a shadow, highlight, or a


combination of these, and can thus easily be detected (see figure 12.28).
We may conclude that planes should be conserved as a family by congru-
ences and similarities, though not necessarily on an individual basis. Of
course the congruences have to respect the structure of the picture as a
planar surface too; thus the “pixels” or “visual rays” have to be conserved
on an individual basis (see figure 12.29). It is remarkable that these gener-
ic constraints fully determine the group of congruences. No additional
arguments are required. When {x, y} denotes the picture plane Cartesian
coordinates, and z the depth, then the similarities are
x5; e µ (cos ◊x : sin ◊y)+ rx ,
y5; e µ (sin ◊x + cos ◊y)+ ry ,
z5; qx x+ qy y + ehz+ e,
282 Jan J. Koenderink and Andrea J. van Doorn

depth

nl
p
np

P
Figure 12.29
“Normal lines” (denoted nl, these correspond to the pixels, e.g., p in the picture
plane P) and “normal planes” (denoted np, these are straight lines in the picture
plane P) are invariants of pictorial space.

where we have included arbitrary Euclidian similarities of the picture


plane for convenience. The Euclidian similarities of the picture plane are
a homothety of magnification exp( µ), a rotation over an angle ◊, and a
translation by rx in the x and ry in the y directions. This Euclidean simi-
larity (or Euclidean motion in the case µ ; 0) is augmented by a number
of true pictorial space transformations. These are a shift e in depth (which
is really irrelevant because it can’t be addressed empirically), a depth
magnification of the Hildebrand type exp(h), and a shear about a plane
containing the visual direction (qx, qy).
Notice that you obtain the identity transform when you set all parame-
ters to zero. The concatenation of any two of these transformations is
equivalent to a third of the same. For any of these transformations, you
can find another one that “undoes” it. Thus these transformations consti-
tute a group.
This is an eight-parameter group. For µ; h ; 0 we obtain a six-
parameter group of motions or “congruences.” When we ignore the
Euclidian similarities in the picture plane (µ ; ◊ ; rx ; ry ; 0), we obtain
a four-dimensional group of essential pictorial similarities. This is the
group that should replace the one suggested by Hildebrand. You easily
verify the group properties in all cases.
Notice that the coefficient h describes the Hildebrand-type depth scal-
ings, whereas the coefficients qx, qy describe the affinities we described ear-
lier (figure 12.30). These are of quite different types, the former being a
similarity, the latter a congruence.
In figure 12.31 we illustrate some congruent objects in pictorial space.
Notice that the shears (“rotations” in pictorial space, shears in the Euclid-
283 Pictorial Space

Figure 12.30
The group of congruences in pictorial space allows the observer to adjust the
“direction of view of the mental eye” ad libitum. Any patch of pictorial relief can
be made to appear frontoparallel.

Figure 12.31
Illustration of pictorial congruences.

ian sense) have effects that are in many respects reminiscent of the trusty
rotations from Euclidean space. They are indeed very similar when small
but appear vastly different when large.

Features as Invariants under Congruency


The group of congruences (the equations listed in the previous section)
defines the geometry of pictorial space in the sense of Felix Klein’s famous
“Erlanger Programm” (Klein 1893). In fact, it turns out that the geome-
try of pictorial space is exactly one of the twenty-seven three-dimensional
Cayley-Klein geometries (Yaglom 1988).
284 Jan J. Koenderink and Andrea J. van Doorn

The relation between two points {x1, y1, z1} and {x2, y 2, z2} that is con-
served by the congruences is U(x1 : x2)2 +(y 1 : y 2)2, that is, simply the
Euclidian distance in the picture plane. One says that pictorial space has
a “degenerate metric” because the metric “forgets” the depth dimension.
When two points coincide in the picture plane they can still be different
because they may differ in depth. For such points the depth difference
z1 :z2 is conserved, and it can be used as a “special distance.” For a pair
of generic points the depth difference is typically not conserved. Thus we
need two types of distances in pictorial space (Strubecker 1956). One is the
distance in the picture plane: it applies to any pair of points seen in differ-
ent directions. The other is the depth difference: it applies (only) to points
seen in the same direction. Both distances are conserved under arbitrary
pictorial congruences.
The congruences (by design) conserve a family of parallel directions
(the direction of the z-axis, or “visual direction”). This conservation has
the immediate consequence that you can’t make a full turn in pictorial
space, except about a visual axis. These latter rotations are trivially the
rotations in the picture plane. If we define proper lines as any lines other
than a visual direction, and proper planes as any planes not containing
a visual direction, then any (proper) plane or line can be brought into
correspondence with any other (proper) plane or line through a suitable
congruence.
When you consider a point in the picture plane that does not lie on the
(projection of) the line of intersection of two planes, the point in the pic-
ture plane corresponds to a point on each plane in pictorial space. The dis-
tance between these points is their special distance. When you divide it by
the distance of either point to the line of intersection of the planes, you
obtain an invariant. This invariant relation between intersecting planes
may be taken as the “angle” between the planes (Strubecker 1956). This
angle is aperiodic. There exist proper planes that fail to subtend an angle
in the generic sense because they are parallel. Such planes have a well-
defined depth difference, though, which can be taken as a special angle.
There exists a full metric duality between planes and points, angles and
distances in pictorial space. Thus we speak of “parallel points” when their
distance, but not their special distance, is zero. This makes the space in
many respects simpler, though in many respects richer, than Euclidian
space. Because of the aperiodic and identical nature of the metrics, we
have two distinct types of similarities in pictorial space. One type (µ P0,
h; 0) scales distances and conserves angles; the other type (µ; 0, h P0)
285 Pictorial Space

Figure 12.32
A cyclical rotation in a normal plane (left). Compare the rotated copies of the
scribble: it doesn’t “turn around,” yet the elements of the scribble change orienta-
tion. On the right a “wagon wheel” in the normal plane. The “spokes” are normal
lines and thus invariant. (The “hub” is at infinity.) The parabolas are “circles” in
this geometry; a rotation shifts them inside themselves, like the rim of a wagon
wheel. The “circles” drawn here are concentric, though they may look “shifted” to
your Euclidian eye.

scales angles and conserves distances. The latter type is exactly the group
identified by Hildebrand. A “general” similarity (neither µ ; 0, nor h; 0)
is characterized through a pair of magnifications, one for the lengths and
one for the angles (Yaglom 1979).
Notice that the “depth shears” (the q parameters) take over the place of
rotations in the normal planes (figures 12.32 through 12.34). They supple-
ment the Euclidian rotations in the frontoparallel planes. With such a
shear you may turn any proper plane into a frontoparallel attitude. Yet
the angle between any pair of proper planes is conserved by these shears;
they thus appear indeed as “rigid rotations.” Although you can rotate any
proper plane into any attitude, the rotations don’t let you make a full turn.
Thus you can’t turn the pictorial object that corresponds to the picture of
a person in a ventral view such as to obtain a dorsal view. Full turns can
only be made about a visual ray, that is, in frontoparallel planes. Such
rotations appear as rotations of the picture plane in itself and are thus
trivial with respect to pictorial space.
The differential geometry of curves and surfaces in pictorial space is
much simpler (though certainly not less rich) than that in Euclidian space
(Sachs 1990). We can define curvature in such a way that it is conserved
by arbitrary motions in pictorial space. We can again introduce principle
directions of curvature, curvedness and shape index, ridges, and so forth.
Although such features have rather similar properties as their Euclidian
counterparts, they are not the same as these.
286 Jan J. Koenderink and Andrea J. van Doorn

+2 p

+1
b
0

q
-1

-2

Figure 12.33
A “protractor” (left) in a normal plane. The shaded sector subtends an angle of
2.25 (simply read the scale). For easier reading you may “turn” the protractor to
bring the bottom line of the sector in a horizontal position (figure 12.34, left,
shows you how). On the right we show the construction of a bisectrix (b bisects the
angle subtended by p and q).

-2

-4 -2 0 2 4

Figure 12.34
On the left, we show the effect of “rotating a protractor.” On the right, a family of
congruent curves. These curves can be transformed into each other via rotations.
287 Pictorial Space

The differential geometric properties of pictorial objects are much more


relevant for perception than their Euclidian counterparts. The latter only
play a role when veridicality is an issue, but as we have seen a great many
important questions can be approached without ever worrying about the
veridicality issue at all, which is a can of worms to begin with and better
left alone when at all possible. Thus the relevant differential geometric
properties of pictorial objects are the “new” ones. This opens up a whole
novel area of investigation in the study of pictorial shape. Such properties
as “curvature” finally acquire a meaning outside the straightjacket of
veridicality. The present theory leads to numerous predictions that await
empirical study.

Multiple Pictures
The police would never record the scene of a crime via a single photo-
graph. It is well known that multiple photographs, taken from different
vantage points, are necessary to produce a reasonably “complete cover-
age” of a scene. This is what most people have done when they bring home
their holiday snapshots. In the limit you would have many pictures in an
almost continuous sequence, that is to say, a movie sequence. The lower
limit is a mere pair of pictures (Koenderink et al. 1997).
It is easily possible to study the type of information in a pair of pho-
tographs that is not available from a single one. We have devised the
“method of corresponding points” to do exactly that. The paradigm is a
most simple one. We present two pictures next to each other and put a
mark on one of them. An observer is asked to put a mark on the other pic-
ture such as to indicate the point in pictorial space that corresponds to the
first mark (see figures 12.35 and 12.36).
The paradigm is perhaps deceptively simple. Notice that the task requires
that the observer has to be simultaneously aware of two distinct pictorial
spaces, one for each photograph. The task cannot possibly be done on the
basis of mere (two-dimensional) image structure (see figure 12.36), unless
there are small local landmarks that show up in both pictures. In our
examples, such landmarks are virtually absent though. The image struc-
tures in the neighborhoods of corresponding points may indeed be com-
pletely different. This task cannot be done by any algorithm as available
today, because it requires the construction of two three-dimensional inter-
pretations each based on a single image. Successful three-dimensional
interpretion in computer vision is still based on multiple image corre-
spondences, though (Faugeras 1993).
288 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.35
Three pictures depicting the same object in the same light field, photographed
from the same vantage point but rotated about the vertical between exposures (by
45° either way). Can the reader find the corresponding points for the black mark
in the center figure on the figures left and right?

Figure 12.36
Two local environments of corresponding points. Notice that the relation to the
contour, the contour itself, and the shading patterns are quite different. It is
unlikely that the task would be possible at all if you only had such a local neigh-
borhood. Observers no doubt use global information as context.
289 Pictorial Space

Figure 12.37
(Left) Result of a correspondences session. (Center) Profile view of a correspon-
dences result. (Right) Covariance ellipses indicating scatter in correspondences.

We find that human observers don’t complain at all when confronted


with such a task but are able to establish correspondences easily (many
hundreds within an hour). From the results we can again construct the
pictorial relief (using standard techniques established in computer vision—
once the correspondences have been established, all the rest is simple), this
time as it pertains to multiple pictures (see figures 12.37 and 12.38). From
two pictures plus a known rotation, we can find the relief. Likewise, from
three or more pictures we can find both the rotation angles and the relief.
The “precision” by which observers are able to do the task depends
on the difference (mutual attitude difference of the objects in the scene)
between the pictures as well as on the nature of the local structure of the
pictorial relief. In figure 12.39 we show an example of the scatter. Notice
that it is as large in the vertical as in the horizontal directions, though the
rotation of the objects was about the vertical. Apparently observers don’t
use such knowledge (obvious from cursory examination of the two pic-
tures) at all. They don’t go by two dimensional cues but perform the task
in the pictorial domains.
Human observers apparently succeed in this intrinsically very difficult
task because they are able to compare the local pictorial shapes of two
pictorial reliefs. From our other experiences with pictorial space, we
have to assume that both pictorial reliefs are only determined up to some
290
Jan J. Koenderink and Andrea J. van Doorn
11.25 22.5 45 67.5 90

500
500 500 500 500

400
400 400 400 400

300 300 300 300 300

200 200 200 200 200

100 100 100 100 100

0 0 0 0 0
0 50 100 150 200 0 50 100 150 200 0 50 100 150 200 50 100 150 200 250 150 200 250 300

Figure 12.38
Profiles of pictorial reliefs obtained from correspondence settings for various rotations between the two images. For the sake of ref-
erence: the rotation due to ocular disparity (vergence angle) would only be a few degrees in this case.
291 Pictorial Space

10 pixels

Figure 12.39
Scatter in correspondence settings. Notice that all disparities are actually limited
to the horizontal direction, but the scatter is roughly isotropic. Observers search
for correspondences in the pictorial spaces, both in the horizontal and vertical
directions.

arbitrary pictorial congruence, most likely different for each. Thus the
correspondences must be due to a comparison of local pictorial shape,
where “shape” is defined as invariants under pictorial congruence.
This use of comparison was studied in a setting where we had three pho-
tographs of a scene containing a mannequin (figure 12.40). The difference
between the three photographs was that the mannequin was rotated by
30° about the vertical between exposures. Because the viewing and light-
ing geometries remain unchanged, both the perspective and the chiaroscuro
of the mannequin in the pictures changed. In this study we did both cor-
respondence settings on pairs of pictures and gauge figure settings on the
individual pictures. We can define relief both from the correspondences
and from the gauge figure settings. We find excellent agreement in the sense
that the reliefs are very similar up to a 30° (Euclidian!) rotation about the
vertical if we first apply certain pictorial congruences (non-Euclidian!) to
the individual reliefs (van Doorn, Koenderink, and de Ridder, in 2001).
(See figure 12.41.) These results show that the observers apparently estab-
lish the correspondences on the basis of the pictorial shapes.

12.5 THE EYES KEEP STARING AT YOU

It is a familiar phenomenon—often enough remarked upon spontaneously


by naive observers—that the eyes of the person in a portrait painted en
292 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.40
For the pictures on the left and right of the center the object was rotated by 30°
(either way) about the vertical. Notice that the light field in the scene was the same;
thus shading and geometry are different for these three stimuli.

face appear to “follow you around the room” (Busey, Brady, and Cutting
1990; Sedgewick 1991; Gombrich 1959). This is typical of pictorial relief,
for the effect fails to occur for a marble bust, for example. The same effect
occurs for people and some objects “pointing out of the picture” (see
figures 12.42 and 12.43).
Here we have a very complicated situation in which at least three dis-
tinct spatial entities play a key role, to wit:
1. The physical space contains both the observer and the picture.
2. The picture frame exists (as a physical object) in physical space, as a
visual object in optical space, and as a pictorial object in pictorial space,
all simultaneously.
3. The person depicted in the picture exists in pictorial space.
Although physical, optical, and pictorial space are distinct, (only) the
picture frame exists in all of them. It tends to appear as a frontoparallel
layer in pictorial space, but “mental eye movements” may freely skew it.
Thus the observer has the freedom to skew the frame in pictorial space
such that it “coincides” with the picture frame in optical space, thereby
removing possibly awkward ambiguities. Notice that this has the effect
of effectively connecting pictorial and optical space in the observer’s con-
sciousness. In such a case the frame would establish a kind of “wormhole”
293 Pictorial Space

100 100

50 50

0 0

– 50 – 50

– 100 – 100
– 100 – 50 0 50 100 – 100 – 50 0 50 100

100

50
50

0
0

– 50 – 50

– 100 – 100
– 100 – 50 0 50 100 – 100 – 50 0 50

Figure 12.41
Comparison of pictorial reliefs from attitude samples and correspondences. The
two rows illustrate opposite rotations of 30° about the vertical. In the left column,
we show a straight scatter plot; in the middle, scatter plots corrected for the best
shear. Clearly the shear component is appreciable here. On the right we show the
residuals after taking rotation and shear into account. Apparently the deviations
are systematic, indicating a piecewise transformation.

between the distinct spaces. What actually happens is of course an empir-


ical issue.
In a recent experiment we used several methods to measure the picto-
rial relief of the picture of a mannequin in a dorsal pose, as well as the atti-
tude of the picture frame in optical space and the apparent frontoparallel
points of the mannequin. We used the picture shown in figure 12.42. We
also prepared a number of projectively deformed copies of this picture. All
pictures were presented on a CRT (cathode-ray tube) screen and were
viewed either frontally or obliquely (at a 45° angle). The retinal image
could correspond to either that in the case of the original picture (figure
12.42) viewed frontally or viewed obliquely (figure 12.43). However, due
294 Jan J. Koenderink and Andrea J. van Doorn

Figure 12.42
A framed picture hung on a wall.

to projective deformation the “oblique” case could also be obtained with


actual frontal viewing and vice versa; thus we decoupled the geometry of
the retinal image from the viewing geometry. Moreover, in using the
gauge figure task the gauge figure could also be projectively deformed (for
instance, in order to counteract the foreshortening effect of oblique view-
ing). We investigated a number of a priori interesting combinations. In
figure 12.44 we show the pictorial reliefs for the cases of frontal and
oblique viewing. Due to the 45° viewing angle in the oblique case, the
width of the relief is severely foreshortened (by a factor of U2/2) in the lat-
ter case. Otherwise the two reliefs are very similar indeed (“same shape”).
Clearly the same features point toward the observer in either case. This is
especially clear from figure 12.45, where we present a scatter plot of the
depths. Any shear component is vanishingly small. There is only a minor
depth contraction. (Depth measured in units of the width, thus corrected
for foreshortening.)
We corroborated this result with an independent method. The observer
had to indicate frontoparallel points (either “near” or “far” points) on
295 Pictorial Space

Figure 12.43
An oblique view of the framed picture hung on a wall (figure 12.42). Notice that
the same features in both pictures appear frontoparallel. This is the same effect as
that of the portrait with “eyes following you.”

Figure 12.44
Pictorial reliefs obtained with frontally (left) and obliquely (right) viewed stimuli.
Due to foreshortening, the right relief is only U2/2 the width of the left one.
296 Jan J. Koenderink and Andrea J. van Doorn

50

– 50

– 100
– 100 – 50 0 50

Figure 12.45
Scatter plot of pictorial depths at corresponding positions for the reliefs shown in
figure 12.44. Although there is a depth contraction (of about 15%), the shear is not
significant.

horizontal levels in the picture. These settings were done in random order.
In figure 12.46 they are seen to line up nicely as ruts and ridges of the
relief. These features are almost identical for frontal and oblique viewing.
The connection of optical and pictorial space is such that (mirabile
dictu) the visual direction may be oblique in physical and optical space
(the wall being seen slanted) and simultaneously in the “straight ahead”
direction in pictorial space (see figure 12.47). Thus the effect of the “eyes
following you around the room” is resolved due to the different angular
metrics in the various spaces (and geometries) involved in the situation. In
retrospect the effect is nothing special.

12.6 PICTORIAL SPACE: WHAT NEXT?

In this chapter we could only scratch the surface of a topic as intricate as


pictorial space. Indeed, much research will be needed in order to arrive at
a somewhat coherent view of the matter. We simply list a few (to our mind
important) topics here.
It is of paramount importance to add to the repertoire of possible oper-
ationalizations and to gain experiences with these. “Good” methods will
feel natural in the sense that the task is immediately obvious and requires
no cogitation to speak of. They yield results very fast, that is to say, at
least a sample of some relevant geometrical quantity every few seconds
for sessions lasting up to an hour. The most useful results are huge data
297 Pictorial Space

Figure 12.46
Frontoparallel points on horizontal scan lines for the reliefs shown in figure 12.44.
Notice that the points line up neatly to form “ruts” and “ridges.” Notice also that
the ruts and ridges are virtually identical for these two reliefs.

ν
Pictorial space Pictorial space
θ
Physical space Physical space
α

n n

Figure 12.47
An observer views a portrait hung on a wall seen frontally (left) and at an oblique
angle (right). The vector n is the normal to the picture plane. In the right-hand pic-
ture the angle b equals 45° (oblique viewing), but the angle o is a right angle! This
is because the line m is the viewing direction, and thus normal to the picture plane.
298 Jan J. Koenderink and Andrea J. van Doorn

sets, of the order of 100 to 10,000 (more is better) samples in some geo-
metrical skeleton framework. Methods are needed that address all the var-
ious aspects of geometry.
It is important to relate the results from as many different operational-
izations as possible to each other. This is the only way to get to grips with
the structure of pictorial space in the generic, abstract sense, detached from
any specific method. Any claims to general conclusions on the basis of any
single method (by no means rare in the literature) should be regarded with
the highest suspicion.
We are still a long way from understanding the nature and efficacy
and—especially—the interrelations of the various cues. The conventional
methods, almost invariably based on stimulus or response reduction, are
highly questionable. We need results in realistic settings (many computer-
graphics-generated stimuli are not above suspicion, because it is rare to
see the relevant physics being taken into account). From experience with
computer vision, we know how difficult and vulnerable interpretations on
the basis of monocular cues really are. The existing algorithms are not
at all robust against small variations on the (typically extremely “ideal”
and thus unlikely) prior assumptions. Thus one should be very wary of
confronting the observer with “cues” that are quite unlike (or even a little
different from) those from the observer’s generic biotope.
We are still a long way from a quantitative theoretical understanding.
This is partly due to the extreme scarcity of empirical data. Most of the
classical data is hardly of the type that could successfully be confronted
with quantitative theories of the type that would enable you to predict
the geometrical properties of pictorial space or objects. It is also due
to the fact that the available data are very hard to detach from the—
often implicit—prior assumptions. In our opinion most of the available
knowledge is a hindrance rather than a help in gaining an understanding,
exactly because it is so intricately wrapped up in unfounded prior assump-
tions. What is sorely needed is a fresh approach, not based on the tangle
of unfortunate terminology that is taken for granted by even most profes-
sionals. This is hard, because who doesn’t like to trade essential uncer-
tainty for the certainty of a word sanctified by silent concensus? For
example, we all nod in understanding when someone mentions “depth,”
yet there may be a dozen (or more) meanings of the term that we do not
know how to relate to one another. Many discussions going on in the lit-
erature are effectively void because one uses the same words for different
entities.
299 Pictorial Space

Pictorial space is quite unlike Euclidian space. We don’t understand it


very well but enough to know that reasoning based on Euclidian precepts
is (at best) unlikely to get us far or—more likely—to lead us into error.
This means that any transformation or operation involved in the compar-
ison of data or the prediction of phenomena on the basis of observed phe-
nomena stands in need of empirical validation. Our Euclidian intuitions
simply cannot be relied upon. This opens up a large field of empirical
inquiry. It is important that such expertise be gained, for further progress
critically depends on it. At this moment we are still at the very beginning.
This page intentionally left blank
Chapter 13
Truth and Meaning in Pictorial Space

Sheena Rogers

Pictures of all kinds fill our world. They hold records of past events, prom-
ises of the future and now, simulations of an alternative present. Some pic-
tures are still, some move, and some surround us, interacting with us in a
virtual reality. All seem to present some truth about the world. Indeed,
perspective pictures, photographs, and movies seem to be fragments of
reality itself, like insects in amber, caught in some medium merely to ensure
their longevity. Seeing the picture and understanding what is depicted
there seem simple. We do not pause to ponder the picture’s success. Young
children follow a story in a picture book before they can read. Wordless
narratives depict biblical events in hundreds of medieval churches. We are
reassured daily of the truthfulness of pictures by their constant presence in
our lives. The pictures we encounter seem to be as simple to see as the real
world on which they present a window.
Photographic pictures seem to capture reality in an especially immedi-
ate way. The medium is transparent. We see through the picture to the
world beyond it (Walton 1984). It has even been argued that “the photo-
graphic image is the object itself” (Bazin 1967, p. 14, emphasis added). So,
when I look at a photograph I literally see my daughter, or a London bus,
the ocean, a chair. “Viewers of photographs are in perceptual contact with
the world” according to Kendall Walton (1984, emphasis added). On this
view, the photographed world is the world itself and so our perception of
the three-dimensional spaces and objects in both must proceed in the same
way. It is a seductive argument, and many writers on picture perception
have fallen for it (see, for example, Walton 1984).
Yet pictures pose a special challenge to theories of perception. Repre-
sentational pictures are flat, yet they depict a world with solidity and form.
Still pictures present frozen moments, torn from the comfortable sup-
porting arms of the spatial and temporal context of real life and held up
302 Sheena Rogers

for our inspection, isolated and alone. Even moving pictures, where action
and observation unfold over time, fix the viewer in space, attached to the
camera, a couch potato unable to step up and engage in the life of the
depicted world (see Rogers in press for more on the perception of movies).
Discrepancies between the location of the observer’s eye and the loca-
tion of the original camera or center of projection generate changes in
the virtual layout of the picture or movie (Rogers 1995). These differences
between pictures and real scenes are important. The possibility of realism
under these conditions must be explained. If there is meaning in pictures,
it does not come easily.
Unquestionably, much of the credit for the success of pictures in carry-
ing meaning must go to the artists. The nature of their contribution
should be examined carefully, however. The devices and designs that pic-
ture makers employ to show a three-dimensional world in frozen frames
have suggested to some that pictures communicate by means of a lan-
guage, an agreed-upon system of signs (e.g., Kepes 1944). Pictorial signs
may “short-circuit” their route to meaning through their visual similar-
ity to the objects depicted (see Metz 1974), but meaning is nevertheless
obtained through a process of decoding. On this view, the image is never
equivalent to the object itself. Realism is an illusion. Although a picture
may tell about reality, it can never be a surrogate for it.
I have presented two extreme views and hinted at decades of argu-
ment about the fundamental nature of pictorial representation. There are,
of course, many variants of these views and many subtleties I have not
expressed. Instead of pursuing these arguments and weighing their relative
merits, I propose that the debate be shifted to empirical grounds. What
exactly do pictures and natural scenes have in common, and can we obtain
experimental evidence of similarities and differences in their perception?
Drawing on the ecological approach to the psychology of perception, I
have argued elsewhere that the perspective structure of pictures can pre-
sent geometrical structures to the eye of the observer that, under certain
constraints, are identical to those available in nature (Rogers 2000). I have
selected one of these structures, the horizon ratio, for empirical investiga-
tion, and will describe it here as a test case in our search for truth and
meaning in pictorial space. The results of the experiments I will describe
here show that although the same geometric structure is available, it is not
always perceived in the same way. Under certain conditions, observers
responded to a pictorial scene exactly as they did to the real scene, and
the picture acted as a reasonable surrogate for reality. The availability of
303 Truth and Meaning in Pictorial Space

equivalent structures does not guarantee perceptual equivalence, how-


ever. Distortions of the perceived space did occur. We will see that there
are limits to the ability of pictures to substitute for real scenes, but under the
right conditions, and in at least some respects, surrogacy can be achieved.

13.1 THE ECOLOGICAL APPROACH TO PERCEPTION

The principal claim of the ecological approach to perception is that there


is meaningful structure in the light reflected from the surfaces of the world
(Gibson 1966, 1979). The available geometric structures carry information
about the nature of events, the properties of objects, and the layout of the
surfaces that make up our world. If an observer can detect these informa-
tive structures, then the world can be known. For pictures to provide the
same kinds of information about the world, light reflected from them must
carry at least some of the same structures. For a picture to be perceptu-
ally equivalent to the corresponding real scene, the same structures must
not only be available; they must be shown to be detected in the same way
and to result in the same perception.
To qualify as informative, in this technical sense, a structure must ful-
fill several conditions. First, it must uniquely and unambiguously specify
some state of affairs in the world under some identifiable set of con-
straints. Second, it must be relevant to some possible action by the per-
ceiver. Third, information “points two ways” (Gibson 1979, p. 141). It
allows us to see ourselves in the world (ibid., p. 225), and in this way it can
guide our actions: “Information to specify the utilities of the environment
is accompanied by information to specify the observer himself . . . to per-
ceive the world is to co-perceive oneself. . . . The awareness of the world
and of one’s complementary relations to the world are not separable”
(ibid., p. 141).

13.2 THE HORIZON RATIO AS INFORMATION FOR SIZE IN REAL


SCENES AND IN PICTURES

The horizon ratio seems to me to exemplify an informative structure. The


optical horizon is implied in the perspective structure of the scene, it is
visible as the origin of optical flow vectors, and it is approximated by the
earth’s horizon. It allows you to see your place in the world because the
horizon itself is always at eye level. As you climb up on a step, or kneel
down, you can see the horizon move up and down in relation to the
304 Sheena Rogers

Figure 13.1
The horizon ratio specifies the size of objects standing on the ground. Objects A
and B each have a horizon ratio of two and are twice eye-height. Object C has a
horizon ratio of four and is four times eye-height. Object C is twice as tall as
objects A and B.

objects standing on the ground around you. The horizon ratio relates your
eye-height to the height of these objects. It specifies precisely the sizes of
the objects relative to your eye-height. In figure 13.1, for example, objects
A and B are each twice eye-height, and object C is four times eye-height.
The horizon ratio is given as the relation between two visual angles,
though it can also be described as a relation of two extents on the pic-
ture plane (Sedgwick 1973, 1980, 1986). In figure 13.2, the angle O sub-
tended by the whole object o from base to top is divided by the angle H
subtended by the distance from the base of the object to the horizon h. The
general version of the ratio is (tanY+ tan H)/tan H, but the angles O and
H themselves approximate the tangents of the angles when they are small,
so o/h / O/H (Sedgwick 1973, 1986).
Objects A and B in figure 13.1 each have a horizon ratio of two
(O/H ; 2). If eye-height is known in some metric, such as feet and inches,
then the absolute sizes of objects in the scene can be labeled in that met-
ric. In addition to information about absolute size based on eye-height,
the sizes of objects in the scene relative to each other are also given by the
305 Truth and Meaning in Pictorial Space

Figure 13.2
The horizon relations are given in terms of visual angle. The horizon ratio is
(tan Y + tan H)/tan H. The horizon distance relation is d ; h ctn H.

horizon ratio. Object C, for example, has a horizon ratio of four and so
it is twice the height of objects A and B. (In the case of both absolute and
relative size, the constraint that objects are standing on the ground plane
must be met. The sizes of objects floating in space, like the classical stim-
uli of early psychology experiments, are not specified by horizon-based
information, e.g., Holway and Boring 1941.)
For our purposes, the horizon ratio is especially interesting. The struc-
ture is present both in real scenes and in all perspective pictures: paintings,
photographs, television, movies, and virtual reality representations. The
optical horizon is always implied in the perspective structure of the pic-
ture, and it may even be visible as a physical line or edge. Indeed, in the
construction of a perspective picture, the horizon may be the first line laid
down on the paper, canvas, or wall. A short stretch of it is often seen in
the background of Renaissance paintings. The horizon defines the limit to
the ground, and in so doing, creates the skeletal structure of a 3-D world.
Although, to my knowledge, horizon ratios have never been used to lay
out the objects in a perspective picture, they could be. Perspective manu-
als usually recommend the construction of vanishing points for parallel
lines lying on the ground plane as the next step (e.g., Ware 1900). Horizon
ratios fall naturally out of perspective structure, however. This is most
easily seen in a picture of a row of objects of the same size receding into
the distance. A perspective line connecting the tops of the objects will meet
a perspective line connecting their bases at a vanishing point on the hori-
zon. The objects, being of the same real size, will be intersected in the same
proportion by the horizon (see figure 13.3).
306 Sheena Rogers

Figure 13.3
The horizon ratio is a component of perspective structure.

The relative sizes of objects (that is the sizes of objects compared to


each other) in the pictured scene are fully specified by the horizon ratio, as
we saw earlier. In pictures, however, an additional constraint is required
to guarantee this information. For the pictorial horizon ratios to match
precisely the horizon ratios from the real scene, the observer’s eyes must
be located in the horizontal plane that extends out of the picture from
the horizon. If the observer’s eyes are lower or higher than this, the ratio
changes. Despite this constraint, the horizon ratio is robust in most ordi-
nary picture-viewing situations. Ratios of visual angles are preserved as
we adopt a viewpoint to the side of the picture, as we walk past it, and as
we approach the picture or step back from it. Whereas some aspects of
pictorial space will deform during these maneuvers, the information for
the relative sizes of depicted objects does not (see Rogers 1995).
A demonstration of the existence of a potentially informative structure
in pictures is not, of course, a demonstration that perception is based on
that structure. Together with my students, I performed a series of experi-
ments designed to investigate the effectiveness of horizon-based informa-
tion for relative size, relative distance, absolute size, and absolute distance
in natural outdoor scenes, and in pictures, movies, and virtual reality. I
will review the findings of this research program before returning to the
307 Truth and Meaning in Pictorial Space

question of truth and meaning in pictorial space and to the possibility that
a picture might act as a surrogate for reality.

13.3 PERCEPTION OF RELATIVE SIZE IN REAL SCENES AND IN


PICTURES

As we have seen, relative size is fully, and precisely, specified in pictures by


the horizon ratio. Using a variety of tasks, participants in my studies have
demonstrated quite accurate perception of relative size in pictures. In one
task participants were shown a simple picture containing only a horizon,
a vertical sticklike object standing on the ground plane, and a small cross.
Starting at the cross, participants drew a line upward until this second
comparison object appeared to have the same real size as the standard. In
experiments using some version of this task (Rogers 1996; Rogers and
Costall 1983) many participants were able to make same-real-size matches
very accurately. There were some limits however—performance was best
when the objects were tall enough to be intersected by the horizon and
when it was clear, from additional information, that the horizon line was
the distant limit of the ground plane.
We have also examined the accuracy of relative size perception in more
elaborate pictures. In two experiments, these are simulations of ground,
sky, and a field of poles generated on a computer in desktop virtual real-
ity (Rogers and Watson 1996a and 1996b). In three other experiments
we used photographs of natural outdoor scenes of a similar field of poles
and compared picture perception to performance in the real scene itself
(Rogers and Pollard 1998; Rogers and Fichandler 2000). In these experi-
ments a set of vertical poles ranging in size from 1 ft. to 22 ft. was arrayed
at distances of 25 ft. to 300 ft. from the observation point. On each trial,
observers made a comparison of the size of two of the poles. They were
told that one pole was ten units tall and asked to make an estimate of the
relative size of the target in the same arbitrary units (a magnitude estima-
tion task). In all studies, whether estimates were made when looking at a
drawing, a photograph, a movie, a drawn scene in interactive desktop vir-
tual reality, or a real scene, relative size perception was quite accurate. In
general, relative size was slightly underestimated at between 95% and 99%
of the size predicted by the horizon ratio. Again there were limits to the
structure’s effectiveness. Accuracy fell when objects were very small, very
large, or when the difference in size between a pair of objects was extreme.
I conclude that the optic array from both pictures and real scenes
contains information for relative size, that this information is based on
308 Sheena Rogers

horizon (or eye-height) relations, and that the information is equally effec-
tive in pictures and in real scenes. The perception of relative size appears
to be achieved in the same way in pictures and real scenes.

13.4 PERCEPTION OF RELATIVE DISTANCE IN REAL SCENES AND IN


PICTURES

Another visual angle relation based on the horizon specifies distance along
the ground to the base of the object by relating the angle of gaze elevation
to eye-height. In figure 13.2, this distance, d, is equal to h ctn H, where h
is the height of the point of observation and H is the visual angle sub-
tended by the extent from the base of the object to the horizon (Sedgwick
1973, 1986). (H is the angle through which one’s gaze would travel if one
were to look at the base of the object and then raise one’s eyes to the hori-
zon.) The angle H becomes smaller the farther the object is from the
observer and the closer it is to the horizon (and regardless of the size of
the object). Observers could estimate the relative distances of two objects
by comparing the angle H for each object. (The angles themselves approx-
imate the cotangent of the angles when the distances are large relative to
eye-height, as they were in my displays.)
We asked observers to make estimates of the relative distance of many
pairs of poles using the magnitude estimation procedure described. In
desktop virtual reality (Rogers and Watson 1996a), in photographs, and
in real outdoor scenes (Rogers and Pollard 1998) estimates of relative dis-
tance were quite accurate. Estimates were between 95% and 100% of the
relative distance predicted by the horizon distance relation.
I conclude that in addition to information for relative size, the optic
array from both pictures and real scenes contains information for rela-
tive distance, this information is also based on horizon (or eye-height)
relations, and the information is equally effective in pictures and in real
scenes. As we found for the perception of relative size, the perception of
relative distance appears to be achieved in the same way in pictures as it
is in real scenes.

13.5 PERCEPTION OF THE ABSOLUTE DIMENSIONS OF PICTORIAL


SPACE

To say that we have perceived the absolute dimensions of a scene is to say


that we know how big or how far something is in some metric. In sharp
309 Truth and Meaning in Pictorial Space

contrast to the discussion of the perception of relative dimensions, the per-


ception of absolute size and absolute distance in pictures is problematic.
Recall that absolute size is given by the horizon ratio in units of eye-height
(as is absolute distance). This fact is still true in pictures, only now, the
eye-height unit refers to the height of the original center of projection
(the station point or spectator point for a perspective painting, the loca-
tion of the camera for a photographic image), which is not known. The
observer’s own eye-height has become detached from the optical horizon
in the pictorial space. Thus the absolute dimensions of pictorial space are
not given. Absolute size and absolute distance are theoretically indetermi-
nate in pictures.
How are picture viewers to understand the scale of a scene shown in a
picture? Are there any circumstances under which a picture could act as a
surrogate for a real scene? There are three possible options: (1) Picture
viewers could scale the pictorial space according to their own eye-height,
even though they cannot know whether their eye-height matches the height
of the picture’s station point. So, for example, if their standing eye-height
is 5 ft. from the ground, by the horizon ratio, objects A and B in figure
13.1 will each appear 10 ft. tall. If they are sitting at an eye-height of 3 ft.,
the objects instead will appear 6 ft. tall. Considerable instability in picto-
rial space would result if observers adopted this strategy. However, if
observers could be anchored at the appropriate eye-height, the pictorial
space should look a lot like the real one it depicts. This artistic (as opposed
to perceptual) strategy is the basis of the stunning 3-D illusions that are
possible in some forms of trompe l’oeil paintings (see Leeman, Elffers, and
Schuyt 1976, for many examples). (2) Picture viewers could use a default
value for eye-height, perhaps their own or a typical standing eye-height.
Margaret Hagen and Robert Giorgi (1993) in fact found that when ob-
servers were asked to locate the position of the camera in a photograph of
a room, they tended to estimate a camera height of about 5 ft. Sitting or
standing, picture viewers would perceive our example objects to be 10 ft.
tall. Of course, the observer still cannot know whether this eye-height
matches the height of the station point, and if it doesn’t, distortions in the
scale of the scene will result. (This is not all bad, as any filmmaker will
attest. A miniature model can stand in for an enormous space ship if it
is filmed from a correspondingly low vantage point. See figure 13.4 for
a related example.) (3) Picture viewers could perceive absolute size as
indeterminate. The detachment of the pictorial horizon from the optical
horizon is visible whenever the observer moves her head because there is
310 Sheena Rogers

Figure 13.4
When we relate our own standing eye-height to the horizon ratio in a photograph
with an unusual camera location, distortions of perceived size occur.

no consequent optical flow within the array from the picture. Instead, it
becomes immediately apparent that the pictorial horizon is not coincid-
ent with one’s own optical horizon. When we do not know the size of the
eye-height unit in the picture, the scale of the space is unknown. Artists
can take advantage of this ambiguity of course. René Magritte, in The
Listening Room (1952), fills a room with an apple. The horizon is clearly
halfway up the apple and halfway up the room, so both have a horizon
ratio of two. We don’t know if we see a large apple in a normal room
(from a standing eye-height) or if we see a normal apple in a dollhouse
room (while laying on the ground). The pleasure of the picture is in the
perceptual uncertainty (see figure 13.5 and plate 4).
The foregoing options are not exclusive. Any one may apply in some
particular circumstance, and it may be possible to create particular per-
ceptual effects by forcing perceivers to adopt one option over another.
Fortunately, the question of how perceivers understand the scale of picto-
rial space is open to empirical testing. We can assess the similarity between
perception of absolute size and distance in real scenes and in photo-
graphs, and we can explore the perceptual options described through care-
311 Truth and Meaning in Pictorial Space

Figure 13.5
The pleasure of Rene Magritte’s The listening room/La chambre d’écoute, 1952, is
in the uncertainty of its reality through the ambiguity of perceived size. Oil on can-
vas, 45 cm _54.7 cm, private collection, © A.D.A.P.G. Paris. © 2002 C. Hersovici,
Brussels / Artists Rights Society (ARS), New York. See plate 4 for color version.

ful manipulation of observers’ eye-height, camera elevation, and viewing


conditions.

13.6 PERCEPTION OF ABSOLUTE SIZE IN REAL SCENES AND IN


PHOTOGRAPHS

Observers in one of the open field and pictorial space studies described
earlier were also asked to estimate the absolute sizes and distances, in feet,
of the poles in the field (Rogers and Pollard 1998). In the real outdoor
scenes, absolute size estimates of 4 to 12 ft. poles, positioned at distances
from 100 to 300 ft., were extremely accurate (close to 100% of the actual
sizes). Observers viewed photographs of the scene on a large-screen com-
puter display with the horizon at the observer’s standing eye-height, and
all visual angles matched to the real scene. Under these conditions esti-
mates of absolute size were near perfect at distances below 200 ft., and
slightly overestimated at distances beyond this.
Information for absolute size seems to have been available both in pic-
tures and in real scenes for the observers in our study. The experience of
312 Sheena Rogers

looking at the large-screen display felt something like looking through a


window at the scene beyond. If our observers were relating the real scene
to their the eye-height in the outdoor study, the conditions we created for
picture viewing probably encouraged the same perceptual strategy they
would use outside in the real world to be used with the pictures.
Corroborating evidence for this suggestion can be found in the same
study. We asked participants in our study for their height in feet and
inches. If one’s own eye-height is involved in the perception of absolute
size as option 1 from the previous section suggests, then there should be a
distinct difference between perception of this property in a real scene com-
pared to its perception in a picture taken from a fixed eye-height. In the
real scene, the horizon will intersect the poles at the observer’s own eye-
height. The resulting ratio will be different for a tall person than for a
small person, but the perceived size of the object should be about the same
for both because the ratio is scaled with correspondingly different eye-
height values. In a photograph, however, the horizon ratio is set by the
height of the camera at the time the photograph was taken. A short per-
son seeing this picture and scaling the horizon ratio with a smaller eye-
height value should perceive the objects to be shorter than would a tall
person scaling the same horizon ratio with a larger eye-height value. (For
a simple example, an object with a horizon ratio of two in the photograph
should appear to be 8 ft. tall to a person with a 4 ft. eye-height but would
appear 12 ft. tall to a person with a 6 ft. eye-height.) And, indeed, in our
experiment, we found a significant tendency for pictured objects to appear
taller to tall observers and shorter to short observers in the photographs
but not in the real scenes. The pattern was not completely clear and there
was considerable variability between observers, but the finding is consis-
tent with the idea that absolute size is perceived in units of eye-height. In
contrast, there was no relationship between the height of the observer and
relative size perception. This suggests that the observer’s eye-height is not
related by the perceiver to the horizon relations that specify relative size.
For this scene property, horizon relations seem to be utilized extrinsically.
I conclude from these findings that the optic array from both pictures
and real scenes provides information for absolute size, that this informa-
tion is based on intrinsic eye-height scaling of horizon relations, and that
the information can be equally effective in pictures and in real scenes.
Perception of absolute size can be achieved in the same way, by the same
perceptual process, in pictures as in real scenes. That is not to say that
313 Truth and Meaning in Pictorial Space

absolute size in pictures is always perceived this way, as we will see after a
summary of our findings for perceived absolute distance.

13.7 PERCEPTION OF ABSOLUTE DISTANCE IN REAL SCENES AND IN


PHOTOGRAPHS

It has long been known that distance is often substantially underesti-


mated, as you can observe for yourself by setting out to walk to an unfa-
miliar distant landmark. It has also been reported that distance in pictures
is usually compressed (see Rogers 1995 for a review). Few studies have
made a direct comparison between perceived absolute distance in real
scenes and in pictures, however. In our study (Rogers and Pollard 1998)
observers looked at a photograph of the real scene displayed on a large-
screen monitor from the pictorial station point; thus the two conditions
are geometrically equivalent. In the real scene, observers underestimated
the distance to the poles quite substantially: at all distances from 100 ft.
to 300 ft. they reported that distances were only about 40% of the actual
distance. In the photographs, distance was compressed to only half this.
Thus a pole actually 200 ft. away appeared to be about 80 ft. away in the
real scene and only about 40 ft. away in the photograph. Here is one scene
property that is not well captured in a picture, even one presented under
conditions that led to accurate real-scene-like size perception.
Distance compression in pictures might be expected under ordinary
viewing conditions. We have two eyes open and can move our heads and
so binocular disparity and motion parallax reveal the smooth flatness
of the picture surface. These sources of information for flatness conflict
with distance information from within the picture itself. Removing them
through peephole viewing has been shown to “undo” the compression and
produce excellent illusions of 3-D space (see Rogers 1995 for reviews).
You can roll up a piece of paper and inspect a nearby photograph through
the tube to re-create the effect for yourself. You don’t even need to be at
the picture’s station point for the illusion to work.
I conclude from this finding that although the optic array from both
pictures and real scenes can potentially provide information for absolute
distance based on horizon relations, this scene property is not accurately
perceived in either case. The truth might be told in the visual angle rela-
tions of the optic array, but we do not seem able to hear it. In this one
respect, pictures are deprived of meaning by conflicting flatness cues and
314 Sheena Rogers

so picture perception is even less truthful than perception of the real scene.
This is an important difference between pictures and life, one that should
be expected in all ordinary picture-viewing situations.

13.8 INDEPENDENCE OF PERCEIVED SIZE AND PERCEIVED


DISTANCE

In the analysis presented here, size and distance are specified independ-
ently by related but distinct horizon relations. Traditional accounts of
size perception, on the other hand, have always held that perceived size
is dependent on perceived distance (the so-called size-distance invariance
hypothesis, e.g., Ittleson 1960). Size, it is claimed, cannot be perceived
directly but is instead obtained indirectly by scaling the single visual angle
subtended by the object with its perceived distance. Thus the causal chain
of perception builds one percept upon another (called percept-percept
coupling; see Epstein 1982). The results of the experiments I have described
indicate that size and distance are, indeed, perceived independently, in line
with the hypothesis that perception of these properties is based on horizon
relations. This finding supports the idea that perception of size is achieved
directly (see Epstein and Rogers, in press, for discussion of other appar-
ently coupled percepts). We found, for example, that despite tremendous
compression in perceived pictorial distance, perceived size remained sta-
ble and was not affected by this distortion. In addition, in the analysis of
individual observers’ data, the accuracy of their size estimates was only
weakly related, if at all, to the accuracy of their distance estimates.

13.9 WHEN IS SEEING A PICTURE LIKE SEEING THE WORLD?

The evidence presented so far strongly suggests that perception of size and
distance is achieved through detection of horizon relations (with the pos-
sible exception of absolute distance). These relations are available in pic-
tures, and the experiments demonstrate that the perception of pictorial
space is governed by the information they provide. Pictures do seem able
to reveal the truth about the spaces they depict (again, with the exception
of absolute distance). We saw, however, that in at least one important
respect, pictures are not reality: the pictorial horizon is not yoked to the
observer’s optical horizon, and therein resides the possibility for pictures
to lie. The height of the point of observation for the picture is not known,
and observers must use one of the solutions described earlier to make
315 Truth and Meaning in Pictorial Space

sense of the pictured space. We have seen some tendency among observers
to use their own eye-height to understand the scale of a pictured scene. If
the location of the original station point (or camera location) for the pic-
tured scene was similar to the picture viewer’s current vantage point, no
harm will be done and the picture will be a truthful depiction of the scene.
But what happens when this constraint is not met? Do we continue to see
ourselves “in” the pictured scene—to see the pictorial space as coextensive
with our own? Can we resist the pull to scale the pictorial world in eye-
height units? Are there some conditions under which seeing the picture is
like seeing the world, and some where it is not?
The pictures used in my experiments so far have all been presented on
a large-screen computer display, with the observers at the station point,
their eye level and optical horizon coincident with the pictorial horizon.
These conditions may have enhanced the observers’ sense of immersion in
the scene through information that the pictured ground plane was co-
extensive with the real one. Observing the photographs this way was like
looking through a window to another place but one still very much a part
of the world we are in, whose ground continued outside the picture to pass
under our own feet. But what of a picture made on paper and hanging on
the wall? And what of a picture made from an unusual spectator point? Is
the perception of these pictures like the perception of the world?
In another experiment, a student and I (Rogers and Fichandler 2000)
manipulated this connection between the observer and the pictorial world.
We compared perception of size and distance in a real scene, with a paper
photograph and with a trompe l’oeil presentation of the photograph in
a peephole viewing box. In addition, we introduced two manipulations
designed to reveal the extent to which observers see themselves to be con-
nected to the pictorial space and so perceive the pictorial space as if it were
reality.
We set up a field of poles as before, on a field next to the ocean with an
unobstructed horizon line. We photographed this scene from four differ-
ent camera elevations: the photographer lay on the ground, sat on a box,
stood on the ground, and stood elevated on the box. We prepared two sets
of these four views as 8˝_10˝ photographs. One set we mounted on the
wall of the laboratory, the other we mounted in a viewing box with a peep-
hole positioned so that the visual angles of the picture matched those of
the real scene. The box was then hung on a structure erected outside
in the actual ocean-side field. The peephole could be positioned at each
of the original observation points for the photographs. Thus, when the
316 Sheena Rogers

pictures were mounted at the correct elevation, they fit smoothly into the
ambient optic array from the real scene. The observer was anchored at the
station point, and the viewing box minimized binocular and motion-based
information for the picture’s flat surface, creating a trompe l’oeil effect.
Under these conditions, the picture came as close to a reality surrogate as
it could. Removing the photograph and flipping open the back wall of the
viewing box gave us the real scene comparison condition, which was
viewed with one eye through the peephole.
The four camera elevations generated four different sets of horizon
ratios for the same scene. If observers use a default value of eye-height,
perhaps their own standing eye-height, as suggested earlier in option 2, the
different horizon ratios in the picture should lead to differences in the per-
ceived sizes of the poles. Objects photographed from low viewpoints will
have larger horizon ratios and should appear larger than objects pho-
tographed from higher viewpoints.
We examined the possibility that as perceivers we relate our current
eye-height, which varies as we move about the world, to the horizon rela-
tions (option 1), rather than a fixed, default value (option 2). Observers
viewed the pictures from each of the four positions, lying down, sitting,
standing, and standing on a box. If they were to ignore their current
eye-height and use the default value, their viewing position should have
no effect on perceived absolute size. If, however, the observer’s current
eye-height is the unit by which horizon relations are understood, then
for a given horizon ratio, changes in eye-height elevation should lead to
changes in perceived absolute size. Lower observer viewpoints should lead
to the perception of shorter objects, and higher viewpoints should lead to
the perception of taller objects.
Observers viewed the four pictures from the one congruent viewing
position for each picture (e.g., sitting photographer with sitting observer)
and from each of the three incongruent ones (in this example, sitting
photographer with lying, standing, and raised observer). If picture view-
ers relate their own current eye-height to the horizon ratio (option 1),
absolute size should be accurately perceived in all congruent conditions,
but distortions of perceived size should occur in the incongruent condi-
tions. For example, the raised observer viewing a picture photographed
from a sitting position should overestimate the size of the object.
The most compelling finding of this study was a difference between the
perception of the trompe l’oeil presentation and the paper photographs.
Observers did not relate their current eye-height to the absolute size of the
317 Truth and Meaning in Pictorial Space

poles in the paper photographs. Estimates of absolute size were very close
to the actual pole sizes and did not vary with viewing position, suggesting
that a default eye-height value was used (option 2). In the trompe l’oeil
condition, however, estimates of absolute size did vary with viewing posi-
tion. When in the lower viewing positions, observers perceived the poles
as slightly shorter than they did when they were at higher viewing posi-
tions. Under- and overestimates of pole size were not as large as would be
predicted by entering the actual value of eye-height into the horizon ratio,
but the difference in the pattern of responding for these two conditions
was clear, and it was statistically significant. I take this difference between
perception of the paper photographs and the trompe l’oeil presentations
to indicate that the observer’s current eye-height can influence perceived
size in pictures (option 1) but only when the observers are forced to treat
the picture as if it were part of their reality. In effect, the trompe l’oeil con-
dition prevents observers from noticing that the pictorial horizon is not in
fact their own. In other words, picture viewers do not necessarily perceive
pictorial space to be coextensive with the space they occupy in the real
world, although under special conditions they can. If they do not, then
there is no reason to scale the pictorial space in units determined by their
actions in the real world.
That is not to say that they do not necessarily bring the same psycho-
logical process to bear on picture perception that they use in everyday
perception of the world. Instead of using the changing eye-height value
generated during movement, picture viewers could still use a default value
of their own standing eye-height (option 2) to extract meaning from hori-
zon relations. Indeed, there is evidence that this is what picture viewers
were doing in our study. Raising the camera made objects appear smaller,
lowering it made objects appear larger, as option 2 predicts. Although cur-
rent actions and body position were not relevant to picture perception,
observers did in some sense see themselves in the pictorial space and see
the pictorial space as part of their world. Once more the observers’ own
body size mattered. Our taller participants saw the poles as taller than
they were, and our shorter participants saw them as shorter than they
were. These two effects occurred in both the paper photographs and in the
trompe l’oeil presentations, implying that the strategy of scaling horizon
ratios with our normal standing eye-height is a general one and not one
specific to picture perception.
I conclude from this last study that picture perception is very like the
perception of real scenes, though there are differences. Horizon relations
318 Sheena Rogers

are the basis of perceived absolute size, relative size, and relative distance
in real scenes and in pictures. (Absolute distance is inaccurately perceived
in real scenes and worse in pictures. It is possible that this scene property
is not governed by available horizon-based information.) Horizon rela-
tions can be exploited extrinsically, that is, without reference to our own
bodies, and this seems to be what we do when we perceive relative size and
distance in both real scenes and pictures. Horizon relations can also be
related, intrinsically, to our bodies. We can use our own eye-height as
the unit for understanding absolute size when it is visible in the optic
array. The evidence of this last study is that at least in picture perception,
eye-height is understood as our usual standing eye-height. We are not
tempted to use a changing eye-height value that results from interacting
with the world (kneeling, climbing, etc.) with ordinary pictures. I would
argue that this is not a “decision” to adopt a picture-specific perceptual
strategy but a behavior guided by optical information. In ordinary picture
viewing there is optical evidence that the pictorial horizon is not linked to
our eye-height. Horizon relations do not change as we approach the pic-
ture and adopt one of the four viewing positions. When they do, as they
would in a real scene or in virtual reality, they reveal our place in the
world, and our momentary eye-height may well become the unit that
scales the dimensions of the pictorial space. When there is no information
that the pictorial horizon is not linked to our bodies (in the trompe l’oeil
condition), we proceed as if it were and relate to the picture as if it were a
real scene. In this case, perceptual judgments are influenced by the posi-
tion of our body and our momentary eye-height, even in pictures.

13.10 TRUTH AND MEANING IN PICTORIAL SPACE

We must return now to the question behind our opening discussion. Do


pictures tell the truth about the world? Do they convey meaning in the
same way as the real scenes they represent? The answer to both questions
is that they can, but they need not.
Pictures can make available to the observer some of the same informa-
tive optical structures that are found in natural optical arrays. However,
a picture can lie in a way that the natural world cannot. When we look at
a picture, we see the picture itself as well as what it depicts. But the picture
is not a layer of glass between a three-dimensional reality and us. We do
not see through a window to another world beyond, though the artist may
319 Truth and Meaning in Pictorial Space

try to convince us that we do. If we are allowed to move our heads, the
optical flow patterns reveal the true nature of the picture as a marked sur-
face. The illusory world will deform and rotate as we move (see Rogers
1995). We cannot always trust the picture to tell us the truth about dis-
tance, or about the sizes of objects depicted.
How, then, are pictures so successful? Why are André Bazin, Kendall
Walton, and others convinced that pictures put them in contact with real-
ity? I claim that it is the very success of most pictures that blinds us to their
origins. Hand your camera to a three-year-old and show her the shutter
release. The randomly frozen optical moments that result in the images
will convince you that meaning is not so easily captured. The absence of
spatial and temporal context, and the absence of artistic control, can strip
meaning from images. To present the illusion of a window on reality, pic-
tures must first capture available informative structures and then support
the meaning in those structures through the satisfaction of the constraints
that underwrite them. Specifically, the horizon relations provide meaning
only when the objects in the scene are on the ground. If the scene is in a
picture, additional constraints apply. For the original horizon relations to
be preserved, the observer must view the picture from within a horizontal
plane extending out of the picture from the horizon. If these horizon rela-
tions are to tell the truth about the dimensions of the scene, the camera or
center of projection must be at a reasonable standing eye-height. Too high
or too low and the picture will lie. Further, if there is any risk that the
picture viewers will not realize they are looking at a picture, they must
observe it from a location that matches their eye-height to the height of
the camera. Too high or too low and the picture will lie.
The potential for distortions and ambiguities in pictorial space have
given fright to many who have pursued a geometrical account of pic-
ture perception like this one (for expanded treatment of this problem, see
Kubovy 1986; Rogers 1995; and Rogers 1997). This potential is handled
here, not by the hypothesis of a shared set of signs that code for distance,
size, and 3-D shape in a picture, or by a cognitive juggling act on the part
of the beholder that fixes any problems, but by the recognition of the
important role of the artist in making pictures meaningful. The careful
deployment of conventional practices (not signs) on the part of the pic-
ture maker can ensure that distortion and ambiguity are minimized, that
the loss of temporal context is compensated for, and that the constraints
under which certain geometrical structures provide meaning are satisfied.
320 Sheena Rogers

The perception of a meaningful and truthful three-dimensional space from


the picture by means of these structures can then proceed in the same
manner as the perception of the three-dimensional world from which the
picture arose. When the constraints are not satisfied, errors in perception
should be expected. The pictured world will not be the same as the real
world it represents. If pictures seem natural it is because they are excep-
tional artifacts. Thus the question of truth and meaning in pictures prop-
erly concerns not if but when they can be found.
Chapter 14
Line and Borders of Surfaces: Grouping and Foreshortening

John M. Kennedy, Igor Juricevic, and Juan Bai

Like an astronomer, a perceiver has the job of using incoming information


to become aware of what sent the information in the first place. The uni-
verse zooms in on the astronomer, and our homely environment projects
information toward us. Then we humble perceivers invert the projection.
This is especially obvious with photographs, because these are flat objects
but we ask what three-dimensional world might have given rise to them.
It is a mystery why a few lines in a brief outline sketch let us readily
perceive figures, faces, and so on. Outlines provide us with precious little
material for constructing a 3-D scene, but we do it anyway. They copy a
few borders and leave the spaces between entirely blank. Their success
tells us boundaries in the scene are vital for much of perception. But, truth
to tell, they are little understood, and the accounts that have been pro-
posed are often quite controversial. That’s why we are investigating bor-
ders here. In a nutshell, much of the story we will tell is that lines (even
dotted lines with prominent gaps between the dots) picture the continuous
borders of 3-D surfaces. In like vein, vivid continuous surface borders
appear in motion and stereo perception, stimulated by elements with gaps
between them. In our account, picture perception, though like real-world
perception, gets remarkable, distinctive work out of its ability to make
groupings, cross gaps, and cause the outlined borders that result to trigger
special geometries of 3-D that are cousins of real-world polar perspective.
All vision ever receives is gradual changes (gradients) in between abrupt
changes (borders). Inverse projection of gradients and borders involves
options, which are useful to outline perception. In a nutshell, inverse pro-
jection of outline provides impressions of surface borders, not all borders,
and continuous borders, even when there is no complete line. This per-
ceived continuity across empty space is the basis for outline depiction.
The perceived continuity used by outline is shared by surface borders.
It is facilitated by evenly distributed elements in close proximity. Outline
322 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.1
Piazza with columns.

uses perspective foreshortening but fails to give us shape-from-shadow


perception. For examples of impressions marshalled by outline, gradients,
and foreshortening consider figures 14.1, 14.2, and 14.3.
Figure 14.1 is a tiled gradient, like a piazza, stretching into the distance.
Two vast and footless columns stand on the piazza, at different distances,
one on the second row of tiles and one on the fourth. The verticals in the
picture that stand for columns bear outline circles, one to our left and one
to our right. The piazza’s rectangular tiles are radically foreshortened by
perspective. The result is we see a level ground (Gibson 1979). Interes-
tingly, we readily take the circles as outlining spheres, and not just flat cir-
cles, but the perspective projections of spheres to left and right of us would
be elliptical (Pirenne 1970; Kubovy 1986). Evidently, impressions given
by inverse projection do not follow the laws of incoming projected light
precisely.
Figure 14.2 is a picture of a 3-D cube. Its bright bars are induced by
gaps between hatched lines in the 2-D picture. We take this to be a picture
of a 3-D cube made of continuous bars. It gives us an impression that the
borders of the hatchmarks are due to continuous edges of bright fore-
ground bars cutting across the hatchmarks.
Further, where four white lines on the page come together to form a
“+,” we can readily see complete horizontal bars in front of partially hid-
den vertical bars. Also, the order of the bars in depth can reverse, so the
vertical bar appears in front. Interestingly, the bar in front can seem to
have complete contours. In the center of the “+,” where the horizontal
323 Line and Borders of Surfaces

Figure 14.2
Cube bars from discontinuous lines.

Figure 14.3
Dotted lines can outline a cube, and a cube’s bars can be shown by solid lines,
dotted strips, and hatched lines.
324 John M. Kennedy, Igor Juricevic, and Juan Bai

and vertical arms come together, the frontal bar can seem to be flanked by
a pair of contours, running across empty space. The pair is horizontal
when the horizontal bar is in front and vertical when the bars reverse in
perception.
Figure 14.3 contains a dotted-line picture of a cube. It also reverses in
depth. The dotted lines function as outlines, suggesting continuous bars,
corners, and obliques oriented in depth. The dots belong together in lines
even though no continuous contour joins them on the page, or in our per-
cept (Arnheim 1974). The dot at the center of the “+” where four lines
come together can be seen as belonging to the foreground bar.
Figure 14.3 also includes a picture of a cube with dotted strips forming
lines, a black line picture of a cube, and a hatched-line picture. Despite
their varied constituents, the cubes appear to have the same proportions.
The obliques recede in depth and are all foreshortened equally.
The dotted strips forming “+” shapes can be grouped as horizontal
bars in front of vertical bars, or the reverse. The central dots of each “+”
belong with the bar in front, in these percepts. They belong to the frontal
bar even with no continuous flanking contour in the percept, much as lines
of dots can group without needing continuous contours.
The black-line cube can seem to have contours crossing the empty space
at the centre of the “+” intersections, much as in figure 14.2’s white cube.
In the hatched figure, some hatches are omitted to distinguish the front
and back bars. The hatches of the front bars are grouped together as a
long continuous row. The line of hatch marks has a discontinuous border
on the page, but the bars being depicted, we have the impression, have
continuous edges.

14.1 OUTLINE AND INDUCTION

Of course, a continuous edge can also be depicted by a continuous bor-


der. Figure 14.4 is a picture of a house, a road, and hills. This little land-
scape was drawn by a blind girl of twelve, Gaia. Blind since birth, Gaia
had some peripheral sensitivity to light for her first six–seven years. She
never had enough vision to see the lines she uses in drawing. She likes to
draw, using a raised-line drawing kit (Kennedy and Merkas 2000). She
has drawn since she was about three years old. Much like other twelve-
year-olds, she draws novel pictures, making up scenes, buildings, people
in intriguing clothing, and so forth. Gaia can feel a line with a fingertip,
one small section at a time, and take it as a long continuous line showing
325 Line and Borders of Surfaces

Figure 14.4
Copy of a raised-line drawing by Gaia. (From Kennedy, in press, with permission
from Prion Press)

a continuous edge, much as sighted people see small sections of line such
as hatch marks and group them as if they had continuous borders.
Gaia’s drawing uses lines for borders of tangible surfaces. The surfaces
of her house are flat, and so their borders are the occluding edges of flat
surfaces. Her hills have brows, that is, occluding boundaries of rounded
surfaces. All surfaces are either flat or curved, so Gaia has deployed lines
to show the borders of the only kinds of surfaces there are.
The skillful use of line in Gaia’s picture invites us to explore the pos-
sible use of outline in pictures. Lines have affinities for tangible borders, in
touch as well as vision, Gaia’s picture suggests (Kennedy 1993). What
kinds of borders can we perceive, besides occluding surfaces, and which
suit outline, and why?
Perceptual borders arise from features in the environment, like Gaia’s
surface edges. The inverse projection problem is to decide which feature.
In the light coming to our eye there may be a “contrast” border between
dark and light. But what was its origin? A flat surface’s occluding edge,
such as the roofline of Gaia’s house? Or a rounded surface’s occluding
boundary like the brow of a hill? Or what?
Perception takes in a limited sample of the light projected to it, and,
in inverse projection, gives us awareness of its possible cause. In this
respect, perception is like the person who notes a few blackbirds in a field
326 John M. Kennedy, Igor Juricevic, and Juan Bai

on Monday, Tuesday, and Wednesday mornings and leaps to the conclu-


sion that there are always blackbirds in the field in the morning. This leap
is subject to the problem of induction.
An inductive leap is fallible. An infinite number of possible generaliza-
tions fit a set of observations. The blackbirds may be there only on week-
days. Or only on days people go past the field. Or only until Fall. Or the
birds may actually be light-colored birds, but the fields are covered with
highly reflective snow.
If the borders in light always had hallmarks indicating what envi-
ronmental feature projected them, inverse projection would be fail safe.
Perception could lock on to the distinctive hallmark and register the envi-
ronmental feature. For example, shadows are darker than their illumi-
nated surrounds. But there are few hallmarks as reliable as the darkness
of shadows at borders. As a result, perception has options at borders.
Generally, perception takes in the border and its setting and uses many
features of the setting as indicators of the origin of the border. If only a
few features are available, in a sketch for example, several options are
open. Because lines in pictures have borders, they can offer these options,
and the more schematic the picture the more options perception can exer-
cise. This ability to exercise options, we hypothesize, explains our capac-
ity for outline representation. But to understand just what options outlines
have, we need to consider the induction problem further.

14.2 INDUCTION

Alas, poor observer, finite data fit an infinite number of curves. Any finite
perceptible pattern can be projected by an infinite number of origins. So
what are we to do?
One drastic error in the study of perception was to argue vision only has
brightness and color as its core because we only have luminance and spec-
tral values to work with. In fact there is nothing to stop us having impres-
sions of any kind whatsoever from stimulation X. We see purple, which is
not in the spectrum, when certain combinations of spectral values come to
us. We see continuity where there is only a group of separate marks. We
get spheres wrong in inverse projection. So it may not be entirely surpris-
ing that borders have affinities with some features of an environment but
not with all.
One cannot predict a priori which elements allow which percepts. The
problem of induction at borders has to be given an empirical solution, for
327 Line and Borders of Surfaces

there is no single solution logically. Psychology has the task of uncovering


the empirical solution, and although the variables used for inverse projec-
tion generally make good sense, as they reflect projective geometry, they
contain many quirks important in picture perception.
Reacting to a problem that is logically insoluble, George Berkeley
thought God must help us to perceive in practice. Denis Diderot thought
we must use independent knowledge of the state of affairs. Hermann von
Helmholtz said we must use inferences (using just another word for induc-
tions). These solutions all accept the problem at face value. Immanuel
Kant realized there must be a set of prior conditions (Cazeaux 2002).
James J. Gibson (1979) offered some prior conditions, but the ones he
advanced were simply to do with the environment. But, it is important, in
doing so he rejected the way the problem was framed. He asserted there
must be a lawful environment in which perceptual data are obtained, not
one with no restrictions. He also contended we could successfully differ-
entiate and integrate our percepts from a given starting point as children,
toward a better fit with the environment. Alas, there are an infinite num-
ber of possible differentiations and integrations, for these are just induc-
tions by another name.
We need another insight besides Gibson’s. Just as the world must be
governed by finite rules, we the observers must have a working beginning
to guide us.
By way of analogy, consider induction of units in words. Words are
made of phonemes. Young children find the relevant phonemes for their
native language in the sounds they hear. But there are an infinite number
of ways of cutting up sounds. For example, all the sounds you hear when
you first wake up might be one phoneme. All the sounds you hear at lunch
might be another phoneme. No one uses those kinds of divisions, of
course. But the probability of any two children, unaided, picking up the
same phonemes from a corpus of sounds is zero. Hence children are aided.
What they must do is have a limited set of possible phonemes, and then
the corpus must match some of that set.
The issue then is arriving at a working beginning in vision, as phonemes
are in hearing. One extremely useful starting point is the border. There
is a closed set of perceptible borders, and a fixed set of origins for these
borders, rendering induction’s problem manageable. Once these sets are
appreciated, other factors can be understood—first, elements that group
as borders; second, foreshortening at some orientations; and third, failure
to show shadows.
328 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.5
Kinds of borders and pictures using them.

14.3 KINDS OF BORDERS IN VISION

The light that comes to our eyes has luminance borders, between high
intensity and low intensity. Likewise, between what we informally call
two “colors,” it has what we may call a spectral border. (The luminance
borders are generally more effective in perception.) Figure 14.5 classifies
borders into luminance and spectral, moving and static, monocular and
binocular. It gives examples of pictures that fit these types.
Luminance and spectral borders can be arranged to form dots, lines,
and textures. The elements (dots or lines or other units) in a texture are
arranged with some regularity over a region. Vision groups the elements
in a region as a unit (a gradient like that in figure 14.1 or a simple form
or gestalt-like lines or squares in figure 14.2), despite the units’ being spa-
tially separate. Trees on a snow-covered hillside define its continuous
slope in vision even if each tree is quite separate. Dots, lines, and contin-
uous borders are evident in monocular vision, with our eye in one vantage
point and a static incoming array.
Motion in an array can also provide a monocular border, much like the
border between two regions. If we were standing looking at two distant
Celtic hills covered by yellow gorse bushes and purple heather, and the
hills partially overlapped each other from our vantage point, they might
look like a single slope. But if we move to one side, we will reveal new
bushes on the more distant hill. The occluding brow of the foreground hill
would become evident. The visible brow would be defined by new heather
329 Line and Borders of Surfaces

and gorse texture being added or “accreted” optically. If we move back to


where we started, the texture will be “deleted,” but this too will reveal the
brow of the hill. The motion-defined border given by accretion and dele-
tion of a few distinct bushes against uniform backgrounds can appear as
continuous as a line of dots, and on occasion it can appear as continuous
as a luminance or spectral border.
Binocular vision can provide borders too—stereo borders. These rely
on relations between the luminance or spectral borders in two eyes, as
motion borders are based on two moments in time. With one eye a grass-
covered lawn may seem like a smooth expanse. Binocular vision may
make out a small ridge ahead in the lawn. If we move, the ridge will give
a stereo border and a motion border in correspondence.
Binocular vision can produce an impression of a continuous border,
given a few spatially separate inducers, just as motion can. This purely
stereo border can appear as continuous as a monocular line of dots, and
it can appear as continuous as a luminance or spectral border.
(We could have several motion-defined borders in one eye and a
slightly different set of motion-defined borders in the other eye. The dif-
ference between the two eyes can provide a stereo border, invisible to
either eye alone, rather like the one in the lawn. This is a purely motion-
defined binocular border. Presumably, we could have many of these bor-
ders, too, say defining small squares or small circles. The shapes may fall
into two regions, with a border between a square-texture region and a cir-
cle-texture region. Indeed, these borders could define further shapes, say
diamonds and ellipses, each in their own superregions. We could define
many such regions, and “superborders” between them. And superborders
could form regions with “superduper” borders, and so on. Perhaps some
of these motion and stereo borders can appear continuous.)
In sum, visual borders can be luminance or spectral, monocular or
binocular, and static or motion defined. The borders can be physically
continuous, and appear so, but also they can be matters of grouping,
appearing as a continuous gestalt even though they are based on elements
that are spatially separate. Often all the kinds of borders coincide in vision,
for color, brightness, stereo, and motion can all tell us about a cliff edge,
simultaneously. But they need not. Markings on the hide of an Alpine cow
have luminance and spectral borders but no stereo or motion ones.
Each source of a visual border indicates borders independently of other
sources to some degree (Cabe, Abele, and Kennedy 1999). Presumably this
is why static monocular vision can suggest borders of three-dimensional
330 John M. Kennedy, Igor Juricevic, and Juan Bai

bars, when we look at a picture like figure 14.2, whereas motion and stereo
vision do not (see also Sedgwick, chapter 3 in this volume): they record the
2-D flatness of the page. The bars look complete, but we also see gaps
between marks. Because each kind of border in vision can be independent
of the others, vision accepts two kinds of contradictions readily—3-D ver-
sus 2-D and continuous versus full of gaps.
If borders carried distinctive hallmarks, as shadows do, then perception
would find them unambiguous. But in fact many apparent borders use the
same inputs. Does a luminance border always coincide with the figure-
ground edge of a foreground object? The Alpine cowhide tells us no
(Kennedy 1974, 1993; Peterson 1999), a lesson we should now explore.

Kinds of Visible Borders


The flanks of a zebra can have different shades or surface colors. Its back
may be silhouetted against a ground such as a sandy-colored hillside. It
can cast a shadow onto the ground, and a shiny wet hoof may reflect an
arc of light on the sand. On the zebra’s belly may be attached shadows,
from the sun overhead. The result is reflectance borders (lightness and
color), shadow borders (cast and attached), and illumination borders on
the sand.
Also, as Gaia’s picture showed, the surfaces in a visible scene can be
curved, like hills, or flat, like a roof. They can be occluding, in which
case there is a foreground and a background (or figure-ground). Or they
can form corners, as in figure 14.6, such as concave and convex corners

Figure 14.6
The outline depicts occlusions, corners, cracks, and wires.
331 Line and Borders of Surfaces

(two foreground surfaces meeting, or figure-figure). Two coplanar surfaces


meet with an elongated crack between them, as in two blocks nestling
together in figure 14.6 (also figure-figure). If a single surface is elongated
and very thin it makes a wire or stick (with background on either side, or
ground-ground), as in the twig in figure 14.6 (Cutting and Massironi
1998).
Surface, reflectance, and shadow borders are the worldly origins for
the borders in light. This much is helpful in providing an answer to
induction’s problem. There are two limited sets—the optical borders and
worldly origins of these borders.
The correspondence between the two sets is not a simple one-to-one
matter. Rather it is many to one. The origin of a luminance border could
be any of the worldly borders. It could arise from a reflectance border
on cowhide, a cast shadow border on the ground, or an attached shadow
border on a belly. Also, it could be from a corner or an occluding edge.
For example, the luminance border could be from a light foreground
object silhouetted against a dark wall.
Because some perceptual borders could be inversely projected as any of
several kinds of environmental features, there are several options for a line
in an outline drawing. But the line may have restricted affinities, that is,
only a few actors, albeit ones influenced by instructions in various settings
(texture, gradients, binocular differences, and accretion-deletion).

Information or Affinity
What normally removes ambiguity at a border? One factor is what lies
between borders. The region between borders is textured. Indeed, the sur-
faces we find in our environs are generally textured. That is, color, lumi-
nance, and texture boundaries usually coincide. Recognizing the presence
of textured surfaces may help us understand the function of perceiving
continuity where the elements are separate. Surfaces have continuous
reflectance and continuous borders but separate texture elements.
Helpfully, the separate texture units on a continuous surface are usu-
ally uniform. On a breezy day, lake surfaces have waves (texture units)
very different from the rocky or grassy margins of the lake. Little wonder
vision has the useful ability to see a continuous region with a continuous
border when the input is separate texture units.
The texture gradients (figure 14.1) in a region help us see a single
surface stretching away from our vantage point. Perspective governs the
change in angle subtended by each unit as its distance increases (Sedgwick
332 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.7
The heights and depths of ghostly columns are shown by the scene’s meridians.

1986; Kubovy 1986). Vision detects how the variations fit with perspec-
tive. Also, it can readily detect the number of texture units between us and
the foot of an object until foreshortening reaches limits set by our acuity.
The result is good perception of depth up to the distance where foreshort-
ening crushes texture units to the degree where they are indistinct (twenty
or so units in a flat piazza such as that in figure 14.1).
Figure 14.7 is a version of figure 14.1 with two ghostly columns. The
grid of converging lines tells us the upper column is larger than the lower
one. Both columns stand on one of the converging lines, but the tops are
aligned with quite different lines in the grid. The upper column projects
three converging lines above the line grazed by the top of the lower col-
umn. Where the line grazed by the lower column meets the upper column
indicates the relative heights of the two. The converging meridians of
the sides of the tiles in the scene indicate height as well as the parallel
meridians of the fronts of the tiles indicate depth. Much of inverse projec-
tion involves applying these informative meridian rules to texture units in
visual scenes.
However, outline omits texture, and the uniform texture between the
lines tells us the picture is simply flat. Is perspective irrelevant to outline
as a result? It does not offer large expanses of darkness, unlike shadows.
Is shading irrelevant, as a result? Shortly we will show just how aspects of
perspective and shadows do and do not matter to drawing in outline. Here
we wish to note simply this: an outline drawing omits perspective gradi-
ents of texture, and shading, and offers static flat surfaces. Its binocular
inspection and accretion-deletion simply indicate it is flat. This suggests,
intriguingly, outline operates despite countervailing information. Outline’s
ability to show a surface edge may therefore be independent of informa-
333 Line and Borders of Surfaces

tion for depth, slant, and occlusion. Rather than information of this kind,
inverse projection of an outline may rely on an affinity between lines and
edges of surfaces. The affinity could be due to common features of lines
and surface edges. The general principle here is that perception sidelines
information and may home in on something in common between what we
use to see our normal surroundings and what we use when we depict the
same scene. If so, theory of pictures must reveal the “common features”
and what factors in perception govern their use.
For clarity’s sake, let us note here that people use two forms of repre-
sentation. One form involves fiat or conventions and the other uses com-
mon features.
Words represent by convention mostly. A hedgehog is smaller than a
pig, even though the word “hedgehog” has more letters than “pig.” Words
sometimes use onomatopeia. The word rivulet nicely stands for a little
stream. A babbling brook is onomatopeic for its referent. When words
stray like crooners into sounding like their subject, they offer a glimpse of
the world of representation by common features.
We have probably only the barest understanding of representation by
common features (Cazeaux 2002). John Searle (1993) once wrote that when
we say “Jill is icy” and mean she is unemotional, the connection between
emotion and being chilled is hard to explain. Being unemotional just is,
somehow, being cold. We detect something in common between the two.
But what? Those of us who have a gift for synesthesia will sympathize
with Searle’s puzzlement over common features we cannot make explicit,
ties that cannot speak their name. Astonishingly, George Lakoff and
Mark Johnson (1999) have argued that much of our thought resides
in implicit connections like synesthesia. The connections are said to be
metaphoric, like John Searle’s metaphor for cool Jill, or Shakespeare’s
warmer metaphor “Juliet is the sun.”
Juliet is as pleasant, comforting, attractive, and essential as a summer’s
day. We do not have to go far to find common features to use in a literal
sentence akin to Romeo’s, though our list of features may not have his
flair. Also, most if not all of Lakoff and Johnson’s metaphors can be
shown to have a literal base. But the search for relevant common features
in matters of representation is by no means easy.
The fact that there are common features for two objects does not
mean that one readily stands for another. Think of two objects. They have
in common that you have just thought of them. But the fact that you
thought of them does not mean one stands for the other. Look at two
objects. The fact that you looked at them together does not make one
334 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.8
Two C-shapes facing to the right, with the illusion that the right-hand C looks
bigger than the left-hand one. If each of the two is flipped in place, without
exchanging places, the left-hand C will look bigger. Subjects then report they are
the same size.

stand for another. Consider two objects that are nearly identical: two cars
off a Rover plant in England. They may be almost indistinguishable, but
that does not make one stand for another, Nelson Goodman (1968) noted.
Context shifts the relevance of features. All objects have many common
features. We have to find some common features and then select the one
pertinent at the time. “Man is a wolf” can shift “hunter” to “killer,” for
example. At a dance, we might say, “he is a wolf,” meaning predatory, and
we might be gossiping about some glamorous and assertive patroller of
the dance floor. In war, we might say he is a “wolf” and mean a vicious
killer.
The problem here, at heart, is the induction problem once again. All
objects have an infinite set of properties in common. Unaided, picking the
proper one has a probability of zero. Hence there have to be guides or
restrictions to our choice.
When we consider pictorial representation, we must ask about relevant
matching features. We can ask what the possible set could be and what
factors influence them. This should be a limited set. Out of the limited set,
only some should provide the relevant match.
Pictorial representation has limits that are not open to change. The ref-
erent of a word is arbitrary, but a picture must look like its referent to
some extent (Wollheim 1973).
It may be helpful to distinguish here the perceptual looks of things from
information gained by watching things alter. Two equal C-shapes placed
side by side, like “( (”, both facing the same way, look unequal (figure
14.8).
335 Line and Borders of Surfaces

Figure 14.9
Illusions add effects: Müller-Lyer (top), Ponzo (middle), and combined figure
(bottom).

But if each of the two shapes is turned in place, to face the opposite way,
like ) ), the apparent inequality will reverse. Observers exclaim, “Oh, they
are equal!” The appearance remains unequal, but the information gained
from the alteration is “they are equal.” Pictures are about providing
appearances, not about the information gained from perceptible transfor-
mations. In a phrase inspired by Heinz Wimmer and Josef Perner (1983),
pictures are about what allows false belief: viewing with restrictions. They
are about what arises when static conditions are provided and what then
induces appearance and illusion.
Helpfully, figure 14.9 shows a distinctive feature of inducers of appear-
ance. The top panel is the Müller-Lyer illusion. The upper horizontal line
looks longer than its twin, just below it. In the middle panel, the Ponzo
illusion also makes the upper line look longer than its twin, the lower line.
In the bottom panel, the two illusions are combined. Remarkably, the
effect is additive (Kennedy and Miyoshi 1995).
Appearances are additive. In this respect, appearances are quite differ-
ent from information. Information does not add; it confirms. Each of the
meridians in figure 14.7 confirms what the others tell us about the size and
336 John M. Kennedy, Igor Juricevic, and Juan Bai

distance of the ghostly columns. If five bits of information tell us one line
is twice another, the five tell us “twice” five times. They do not add to tell
us “ten times bigger.” (We should note that if the Ponzo causes a 5% illu-
sion and the Müller-Lyer 10%, their combined effect may be 14%, because
of common factors. If the first illusion is due to factors a, b, and c, and the
second to factors c, d, and e, their combination offers a, b, c, d, and e, with
c as a common factor only appearing once, not twice. The process is addi-
tive, but the sum is an imperfect addition.)
Pictures govern appearance, straying at times into illusion and additive
effects. In particular they offer lines giving the appearance of edges, influ-
enced by perceptual factors given by the patterns of the lines.
Pictures are unlike words, because words, despite onomatopeia, are not
essentially about looks. Also, phonemes, the components of words, do not
mean things, but the looks of pictorial borders may rely on common fea-
tures with other borders. So let us ask about the features in pictures such
as Gaia’s line drawing.

14.4 NUMBER OF BORDERS

One simple theory is that for each border in the depicted world there is
one in the picture, and no more. This is hardly true for outline, a moment’s
thought reveals. An abrupt border on a surface we can call a contour, an
abrupt border in the light to the eye we can call a contrast, and an abrupt
change in depth at the border of a flat surface we can call an edge. Inter-
estingly, pictures such as Gaia’s have lines formed of two contours close
together, forming a line in an outline picture. Although the outline in the
picture has two contours projecting two contrasts, it generally is taken to
stand for one edge—a single edge of a single surface (Kennedy 1997). That
is, the referent of two contours is one edge of a surface. The affinity of line
for edges is independent of number of borders.

14.5 CURVED AND STRAIGHT LINES AND CURVED AND STRAIGHT


SURFACES

The second explanation we might consider is one I owe to Nelson Good-


man (personal communication, 1972). I pointed out to him that lines can
show edges of flat objects and boundaries of curved objects. He wondered
if the reason was lines are straight and curved, and their referents are
straight and curved. The common property is shape.
337 Line and Borders of Surfaces

This theory explains too little (as Goodman’s Project Zero colleague
David Perkins pointed out at the time). Certainly common shape proper-
ties are important. But there is more. Edges of flat-surfaced objects such
as cubes are indeed depicted by straight lines. But the boundaries of cylin-
ders and water glasses are too, and these are curved. We see the straight
lines stand for boundaries of surfaces curving away from us. The affinity
of lines for edges does not depend on matching straight for straight and
curved for curved (Willats 1997). If a close match of shape were vital,
young children could not draw in outline.

14.6 JUNCTIONS OF LINES MATCH VERTICES OF OBJECTS

A third explanation to consider is based on the vertices or junctions of


lines. These form shapes such as L-junctions, T-junctions, Y-junctions,
and X-junctions, and at ends of lines there are simply termini (abrupt or
fading). When we draw a cube, we draw vertices as well as straight lines.
Junctions have common shapes with vertices of cubes. Does that explain
the success of the outline drawing? Gibson (1979) thought so.
This explanation also falls short. The explanation emphasizes junctions,
which do indeed look like vertices of cubes, but it neglects the lines join-
ing the junctions. The lines between the vertices look like edges. At the
least, the junctions have to extend an influence to the adjoining lines. In
addition, proportions of regions between junctions are important. At one
extreme, using a kind of connectionist topology, the argument could be
that the set of vertices are arranged in a certain order and that is all that
is important. But if only topology mattered, squishy puddles of orderly
vertices would be identified as cubes. In fact, we need more. There are
definite effects of proportions.

14.7 CONCENTRATION OF ELEMENTS

A variant on the explanation emphasizing vertices is a theory speculating


that perceptual information is concentrated at vertices and other points of
rapid change (Attneave 1954; Biederman 1985; Kennedy and Domander
1985). Alas, experiments testing this theory were confounded. In one con-
dition, only regions between vertices were removed. In another, only ver-
tices were removed, supposedly, but in fact vertices and regions between
vertices were both removed. If the parts are removed in the same fashion,
there is little difference between the two sets of pictures (figure 14.10), and
338 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.10
Cups with partial deletion of outline. Omitting sections between vertices (left)
creates the same problems for perception of the cups as omitting sections with ver-
tices (right).

some parts of the figure with vertices removed, such as handles, are easier
to perceive. Further, five evenly distributed dots on each line in a picture
of a box suggest the overall figure more readily than dots concentrated
midway between vertices, and least effective is one with dots concentrated
at vertices (figure 14.11). If the figure has curved lines, like a bird does, one
with dashes evenly distributed suggests the overall form much more effec-
tively than one with dashes concentrated at changes of direction (figure
14.12, from my Toronto student Ramona Domander). A figure with many
indentations, such as a hand, also tells the story. The hand with evenly dis-
tributed dots is easier to see than one with dots concentrated at points of
change (figure 14.13, from my Salzburg student Christoph Obermair).
Figure 14.13 includes two hands made by joining the dots with straight
lines. Instructively, the one made from the concentrated-dots figure looks
more like a hand. Evidently, when we look at the even-dot figure we do
339 Line and Borders of Surfaces

Figure 14.11
Boxes and dotted lines. The box shown by lines with evenly spaced dots (bottom
right) is easier to see than the box with dots concentrated at midlines (top right),
and the poorest figure is the one with dots concentrated at vertices (center left).

not just join the dots with straight lines. We fit curves. Joining the
concentrated-dot figure with straight lines succeeds because then its dense
dot regions approximate curves.
Even distribution is hardly the sole factor grouping elements in lines.
Density matters. Like the bird pictures, the hands use the same number
of elements, distributing them over two perimeters with the same length.
Overall density is constant, and even distribution is shown to be signifi-
cant. However, we group figure 14.14 (from my Toronto student Rizvana
Chishty) as five straight columns of diamonds with irregular spacing. We
340 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.12
The bird figure with lines evenly spaced is easier to see than the bird figure with
lines concentrated where contour changes direction.

cannot readily see figure 14.14’s seven rows of perfectly evenly (horizon-
tally) spaced diamonds, orthogonal to the five irregular columns. Vision
favors the five columns with irregular but denser spacing.

14.8 FORESHORTENING OBLIQUE LINES

The distribution of elements between vertices influences vision. The pro-


portion of these parts is important in outline shape perception. But how?
Children are said to draw what they know, not what they see. In this
popular theory, outline drawings of cubes by children are not based on
perception and inverse projection, using foreshortening, but rather on
drawing lines in the same proportions as the edges of the object.
341 Line and Borders of Surfaces

Figure 14.13
The hand with dots evenly spaced is easier to see than the hand with dots concen-
trated where contour changes direction. But if the dots are connected with straight
lines the one based on concentrated dots is more like a hand.

Consider lines on the page that are oblique, horizontal, or vertical in


figure 14.2. Andrea Nicholls and Kennedy (1995) asked adult subjects
about drawings of blocks, like those in figure 14.2, with short and long
obliques. The subjects were asked which obliques suggested receding lines
equal in length to the square’s lines. Subjects selected oblique lines that
were about 65% of the length of the square’s lines. Further, figure 14.15
shows that blocks drawn using frontal squares and obliques of the same
length are taken as having obliques much longer than the lines in the
squares. The angle of the oblique varies from 25° to 65°, with no change
in this apparent proportion as observed by subjects. Subjects also judge
the obliques on the page to be about one-third longer than their true pro-
portions, when judging obliques 1.5 to 5 times longer than the square’s
side.
342 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.14
Columns of irregularly spaced diamonds are seen, not regularly spaced rows with
less density.

In another study, Nicholls and I changed the frontal square into a tra-
pezium (figure 14.16). This has four equal-length obliques and no right
angles. Now the oblique misjudgments should apply to both the front face
and the long obliques, when subjects assess their length on the page, and
they should be compared more fairly. Indeed they are. At 1 (front trape-
zium) to 1.5 (long oblique) they were judged veridically. Further, given an
oblique of 5 times the front trapezium the judged ratio was 1.1 times the
true ratio.
If the front face of a brick is depicted by a trapezium, the brick appears
to recede at an acute angle to the picture surface. If the front face is a
square, the brick recedes as if orthogonal to the picture surface. The result
is more foreshortening for the square-fronted drawing. Obliques of length
343 Line and Borders of Surfaces

Figure 14.15
The obliques are seen as longer than the horizontals and verticals of the bricks.

Figure 14.16
Trapezium fronts invoke less foreshortening of obliques than do squares.
344 John M. Kennedy, Igor Juricevic, and Juan Bai

1.5 appear longer in the square-fronted drawing than in the trapezium-


fronted one.
Why should oblique lines be judged longer than verticals and horizon-
tals? The answer surely is that foreshortening applies to obliques showing
edges receding from the frontal-parallel picture plane. Why subjects use
65% as the best ratio to perceive 1:1 length is as yet unexplained. (The
actual length on a picture surface of the oblique attached to a square
showing the frontal side of a cube ranges from zero to infinity, in polar
projection. It is zero for a cube in front of the center of projection, small
for a cube slightly to one side, and infinite for a cube far to the side;
Kubovy 1986). We do not know what the preferred ratio is for many of
these cubes. And we do not know what the ratios are for cube drawings
subtending various angles at the eye. The 65% ratio here is just a helpful
starting point for investigation.)

Drawing and Foreshortening


An outline showing a cube stimulates perceptual impressions, not just
preference judgments. Its obliques give misleading impressions of size,
because of foreshortening. However, the theory that children draw what
they know and not what they see predicts that children drawing cubes will
make all lines equal on the page. They know all the sides are equal in the
cube, so they will make all the lines equal. To put this idea to rest we need
to measure what line lengths children do in fact use. And indeed, children
of nine and ten use foreshortening in their drawings of cubes (Nicholls
and Kennedy 1992, 1995). They do not use a 1:1 ratios of frontal square
to oblique. Rather they use 60%–70% ratios, foreshortening in a fashion
that fits the foreshortening ratios used by vision.
Foreshortening is inverse projection. It influences perception in many
ways, though its effects cannot be predicted a priori. Indeed, if we set
out to draw pictures like figure 14.1, we often fall into very sizable error.
We are often dramatically wrong when we try to estimate the fore-
shortening of objects at different distances, if they are not optically adja-
cent (Kennedy and Juricevic 1999). This is why painters hold up pencils
to compare relative visual sizes of distant objects.
We can demonstrate the painter’s problem as follows: bend your left
arm. Also, put it halfway between straight ahead and directly to the left.
Then pose your left hand close to vertical. Now stretch your right arm far
to the right. Pose your right hand close to vertical. Your arms should be
135° apart. One hand should be twice the distance of the other from your
345 Line and Borders of Surfaces

eyes. Now try to estimate the relative visual angles of the two hands (their
vertical extent). Most people say the right one looks about 90% the visual
angle of the left. Now swing both arms to straight ahead, being sure to
keep one bent and the other straight. Now the right hand’s visual angle
looks to be about 70% of the left’s. Now close one eye. Be sure the hands
are optically adjacent, but the right is still twice the distance of the other.
Now the right’s visual angle looks close to 50% of the left’s. Evidently,
visual angle ratios of the hands are judged to be close to their ratio in
length (in centimeters) when the two hands are set widely apart in angle,
and we view binocularly (getting good length information). Igor Juricevic
and I found these effects in studies with real objects set on a tabletop,
and studies with pictures of objects like that in figure 14.1. Judgments
of visual angle ratios vary from 100% veridical when objects are aligned
(optically adjacent) to judgments that are increasingly like length ratios as
angular separation increases. The judgments move steadily toward length
ratios, and come close to 100% of length ratios when objects are set 180°
apart (one to our extreme left and one to our extreme right).
Little wonder we have trouble drawing in perspective! The relative
visual angles of nearby objects are misperceived quite massively. Photo-
graphs arrange patches on a picture surface to match visual angle ratios
of objects in the 3-D world. To draw with photographic realism, we need
to judge relative visual angle ratios of the 3-D target objects. We cannot
do this accurately without aligning objects optically. That is why artists
hold up pencils optically alongside target objects.
Consider using a pencil like an artist when drawing two identical elec-
tricity pylons, one near and the other far, one to our left and one to our
extreme right. The ratio of the nearby pylon’s visual angle and the pencil’s
are judged correctly: say 0.6 to 1. Then a second pylon is assessed against
the pencil: Say 0.3 to 1. This tells us the ratio of the two targets’ visual
angles is 2 to 1. If we had not used the pencil we might have judged the
visual angle ratio of the pylons as say 1.2 to 1, close to the height ratio.
The effects of foreshortening are related more and more to length and
height information as target objects separate in angle. Thus vision is not
predictable just from the laws of perspective. We should use foreshort-
ening, because it is everpresent, but some impressions we get from it are
not predictable from laws of perspective alone. Pictorial impressions
are to do with these effects, which are dramatic and little understood.
The effects require empirical investigation, not just a priori geometrical
considerations.
346 John M. Kennedy, Igor Juricevic, and Juan Bai

14.9 YOUNG CHILDREN USE OUTLINE

Outline invokes impressions, we have argued. However, in the earliest


years of using lines, say from ages two to five, Constance Milbrath (1998)
has argued, children are not using line to stand for perceived edges of
surfaces and do not make any use of the observer’s vantage point. This
theory may underestimate the child. First use of outline for the visible bor-
ders of a person or an animal occurs at about age two or three. Similarity
of shape is evident in drawings of children of about three (Kennedy,
Nicholls, and Desrochers 1995). These early and rough outlines show bor-
ders of objects. Notably, they show the occluding boundaries of rounded
objects. Hence, they imply a vantage point (Hopkins 2000; Lopes 1997;
Willats 1997).
In a remarkable discussion, Milbrath argued that outline is something
only talented children use especially effectively at young ages: “talented
children quickly caught on to the idea that a line should stand for an
edge and a plane for a surface” (p. 357). This much is surely harmless, and
follows from her observations of two-year-olds and kindergartners. But
Milbrath went on to claim most children do not catch on easily to using
outline effectively (pp. 181–182), and by adolescence many of the “less tal-
ented children” have not done so (p. 357). But surely viewer-centered fea-
tures, implicating the observer’s vantage point, are what any child draws
when using outline at any age. Outline readily shows occluding edges of
surfaces, and these edges of surfaces are surely available early in the devel-
opment of perception.

14.10 LINES AND DISCONTINUITIES

In Gaia’s picture, lines clearly depict surface edges. But should we gener-
alize and suggest lines can stand for any discontinuity? Let us consider
two theories, one arguing that any discontinuity suits outline, the other
arguing that lines show surface edges but not purely visual margins such
as those of a shadow’s.
A line has two contours. A black line on a white page is a thin black
region with white surrounds on both sides. The luminance border that
relies on black giving way to white is the shadow border.
Lines can match the shapes of any spatial discontinuity, including a
border of a shadow (Kennedy 1974). But what can they look like? Inter-
estingly, they cannot look like shadows, or objects bearing shadows.
347 Line and Borders of Surfaces

Figure 14.17
Shape-from-shadow fails in outline because it provides a negative contour along-
side the shadow.

A face in strongly directional illumination can have many shadows, cast


on the cheek by the brows and nose for example, or attached to the side
of the nose. High-contrast pictures of faces with cast and attached shad-
ows work quite well in perception (figure 14.17, leftmost panel, solid black
figure, and rightmost panel, solid gray figure). They trigger a “shape-
from-shadow analysis” that allows us to see the face, its 3-D configura-
tion, and its expression. Such pictures have two tones, one bright for
illuminated regions and one dark for shadow regions, in a “positive.” (A
“negative” reverses the positions of the bright and dark, so the bright
region in the picture stands for what was in shadow in the original.) The
grouping of dark patches in the picture reveals well-defined areas on a
continuous surface with several slopes, showing the face surface rising and
falling over forehead, nose and cheeks.
If outline depicted any spatial discontinuity, and hence any source of
contrast patterns, it should be able to stand in for shadow patterns. We
should be able to use a line in place of the discontinuity between shad-
owed and illuminated regions. The outline should allow perceptual analy-
sis to provide a percept of the face, its configuration in depth, and its
expression. The grouping of the patches surrounded by lines should
seem ordered, and firmly grouped, and not like independent puddles or
islands.
In fact, outline cannot readily stand in for the borders of shadow
patterns (see figure 14.17, outline figure, second from left). It does not trig-
ger shape-from-shadow analysis (Kennedy 1993, 1997; Kennedy and Bai
2000; Cavanagh and Kennedy 2000). Likewise other purely visual borders
such as color borders cannot be replaced by outline with any degree of
success. Just as we do not see outlined shadow patterns look dark, we do
348 John M. Kennedy, Igor Juricevic, and Juan Bai

not see familiar color forms look colored when shown purely in outline.
Hence the discontinuity theory is too general.

14.11 OUTLINE MATCHES ONE CONTOUR

Could the explanation of outline be that a line uses one of its black-white
borders to match the referent’s border? That is, some borders, such as
shadow borders, are “polarized” so that they are darker on one side. A
shadow is darker than the illuminated region neighboring it, a factor that
plays an important role in shadow perception (Cavanagh and Leclerc
1989). Does outline work if it matches a polarized border?
A theorist might contend that an outline will work like a shadow if it
had a contour with the shadow’s polarization. This hypothesis fails, alas,
because one of the black outline’s two contours in figure 14.17 has pre-
cisely this pattern of polarization—the contour alongside the white, illu-
minated region.
Another explanation of the failure of outline to trigger shape-from-
shadow is that outline has two contours, and a shadow border has one.
Alas, a shape-from-shadow figure with two borders succeeds, provided
both are polarized in the correct direction (figure 14.17, second panel from
right). That figure has a black area for shadow, a grey area for a line, and
a white area for illuminated regions. Indeed, in studies with Juan Bai, I
have found that more than two borders can be provided and if they are
all polarized like a positive, shape-from-shadow succeeds. A successful
explanation of the failure of outline with shadow borders is that one of the
line’s contours offers incorrect polarity at the border of the shadowed
region. Hence the failure of a figure with a gray region for shadow and
a black outline (figure 14.17, third from left). The figure has gray for shad-
ow, a black line, and a white area for illuminated regions in the referent.
The line’s contour alongside the illuminated region has the correct polar-
ity. The line is dark and the illuminated area bright. But the contour it
provides to the shadowed region is polarized like a negative. It is gray on
the shadow side and darker on the line’s side.

14.12 OUTLINE AND GROUPING

Up to this point, the list of theories of outline has treated outline as using
continuous contours. But outlines with gaps work well too. We can draw
349 Line and Borders of Surfaces

a cube with lines of dots. If outline can be served by lines with gaps, then
it is the grouping of the elements that is at work. The axis of the grouping
is crucial. This has a location and a direction from us, but no color or
brightness. Hence it does not matter to outline whether the marks on
the page are black or white dots, or dashes or continuous lines, as figures
14.2 and 3 reveal. (I should note that Willats [1997] disagrees with this
conclusion.)
What do lines with continuous and discontinuous borders share with
their referents? The referents are surface borders, that is, corners and occlu-
sions. Any given outline is therefore ambiguous. The pattern in which it
is found helps it take on a referent in perception, but inverse projection
of outline works with vague sketches with no decisive features. The key
therefore is some common feature of a dotted line, with gaps, and a con-
tinuous surface border. But what feature?
Surfaces can be textured sparsely or densely. The contrasts telling us
about borders of surfaces are sometimes continuous optically, but some-
times they are full of gaps. Likewise, random-dot stereograms have regions
with no continuous contour in the stimulus display, but binocularly they
give us an impression of a continuous border. The border has no bright-
ness polarity, just like dotted outline. Similarly, kinetic random textures
use accretion and deletion of elements, with discontinuous stretches where
grouping of elements occurs across gaps, to push us into percepts of sharp
borders. Again, these borders are not brightness-polarized.
In sum, the optical patterns that specify borders of surfaces have fea-
tures that match the patterns operating as outline. Both kinds of patterns
can provide continuous stretches of luminance or spectral contour, but
often they offer gaps that need to be bridged. The physical factors that
permit grouping and apparent continuity, without polarization, can be
triggered by both outline and information for surface edges.

14.13 AMBIGUITY OF CONTOUR AND OPTICAL EDGE INFORMATION

Outline is ambiguous about its referent in the absence of context. So


too is a continuous luminance or spectral contour. It might, however, be
thought that accretion and deletion borders (Gibson 1979) are distinct
from outline in that they are far from ambiguous. Accretion and deletion
might be thought to specify occluding edges without recourse to context.
If this were so, the similarity of outline and accretion-deletion boundaries
350 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.18
Accretion and deletion occurring with moving opaque foreground objects, in uni-
form illumination, with no shadows.

would be threatened. Only outline would need context. However, accretion-


deletion contours are ambiguous about the role of occlusion, both physi-
cally and perceptually.
One well-known physical condition creating accretion and deletion is
an opaque object passing in front of a textured background. The leading
edge of the opaque object deletes background texture and at the trailing
edge background texture accretes. Texture that is accreted and deleted
indicates background (figure 14.18).
However consider the case of snow falling from a white sky against a
dark background (such as distant hills, trees, or windows in white build-
ings) and then falling further against a white background (like a snow-
covered field or a white wall of a building). Often, the snow is only visible
against the dark background. The snow accretes at an upper edge of a
dark patch in the background and deletes at a lower edge. The same is true
for rain. The accreting and deleting elements are in the foreground. The
accretion and deletion does not cause us to see dark trees as holes into a
world beyond.
Similar effects occur for smoke rising, or leaves whirling upward, or
insects swarming. Often they are only visible against a light background.
Also, the wind can create a whorled pattern across a field of wheat, or a
351 Line and Borders of Surfaces

Figure 14.19
Accretion and deletion occurs for snow accreting in a beam from a light. The A
elements are accreting, and F elements are falling. A cross section of the beam has
accretion for elements at the left and right borders. A frontal view also has accre-
tion in the center regions between the left and right borders.

water surface. At the leading edge, optical units accrete, and at the trail-
ing edge units delete, and we see a single continuous surface. The elements
accreting and deleting are on the surface of water and the top layer of
wheat.
Related effects occur at night with directional illumination. As an obsta-
cle moves between us and a light source, its shadow sweeps through the
falling snow. Deletion occurs at the leading edge of the shadow and accre-
tion at its trailing edge. Snow can fall into a light beam, from a dark sky,
and accrete in the optic array. In figure 14.19, the A units are accreting and
the F units are falling. Snow can fall through the horizontal light beam
from a car headlight, deleting optically at the lower edge of the beam.
The particles seen in a street lamp’s light can be seen falling until their
background is the snow-covered sidewalk. The continuing dance down-
ward, visible when passing in front of a dark wall, is no longer visible
against the white sidewalk. The snowflakes accreting and deleting are in
352 John M. Kennedy, Igor Juricevic, and Juan Bai

Figure 14.20
An object racing to the right casts a shadow on elements (E) with deletion (D) at
its leading edge and accretion (A) at its trailing edge. The elements in the shadow
are not illuminated and are invisible.

the foreground, and the wall is the background. At night, insects swarm-
ing in a beam of light accrete and delete at the beam’s margins, contrast-
ing with dark buildings in the background.
In a 1930s gangster movie set in Alcatraz, a spotlight running across a
surface such a brick wall at night could bring elements into the optic array
and then move on so they would fall back into darkness. Bricks accrete
and delete revealing a continuous surface, with no foreground or back-
ground. Cagney or Bogey could run through the spotlight and cast a deep
shadow on the ground that would delete elements at its leading edge, while
elements would accrete at the trailing edge (figure 14.20).
Other cases are provided by motion and highlights. As we move past a
series of small spots of water on a surface, a path of highlights can stretch
from us towards the sun, especially if it is low in the sky. The path moves
with us, staying aligned with the sun. The highlights accrete as we move
forward and delete as we move too far. The sparkling lights accrete at
the leading edge of the path and delete at its trailing edge. The accretion
and deletion tell us about a continuous surface below us. This is vividly
observed flying east-west past the thousands of kettle lakes of the Missis-
sippi, with the sun to the south, or walking over a fresh dry sparkling
Ontario snowfall on a sunny winter’s day.
On a breezy day, lake water often reflects sunlight as rippling light
patterns on boat hulls. The ripples run along the hull and delete at the
surface’s edge. The accretion and deletion tell us about the foreground
surface.
In sum, like outline, accretion and deletion require grouping to tell us
about a continuous border. Also like outline, they are ambiguous and can
353 Line and Borders of Surfaces

use context to tell us to see them with one foreground-background rela-


tion or another.

14.14 STEREO BORDERS

Stereo perception is also very capable of linking separated elements. In


random-dot stereo displays, separated dots trip very effective percepts of
continuous edges. As in misunderstandings of accretion and deletion, it
might be thought that stereo-defined depth is fixed for each element.
Context is unimportant, in this view. If so, the hypothesis that there is
a role for grouping for all edges, and that outline shares this feature, is
threatened.
In fact, stereo-defined elements are affected by context, as a cube based
on figure 14.3’s dotted-strip figure can show. A flat cube drawing can be
defined in a random-dot stereogram. Each stereo-defined line in the draw-
ing is, say, six or seven dots wide. The stereogram can have all the dots
defining the lines be at the same depth. The lines are coplanar but stand
out from the background of random dots. That is, if we only saw small
sections of the lines of the stereo-defined figure all the lines would appear
to be on one flat surface, like a mass of dots on a transparency, above a
flat, dotted background.
The flat stereo-defined drawing does not just stay flat. In inverse pro-
jection, it goes into apparent depth, We see the flat sheet of lines (a stereo
image that looks like a 2-D pattern) and we also see a depicted cube, its
bars in 3-D. Further, at “+” intersections, the apparent depth of the bars
depicted by the stereo-defined lines reverses, much as the depicted bars do
in figures 14.2 and 14.3. That is, at the “+” junctions formed by the lines,
the percept first causes, for example, the horizontal stereo-defined bar to
be in the foreground and continuous, and then the vertical stereo-defined
bar to be in the foreground and continuous. The dots in the center of the
stereo-defined “+” first group with the other dots to left and right in the
horizontal bar, and all these dots are in the foreground. After a reversal,
the central dots group with the dots above and below, in keeping with the
vertical bar, and all these dots are in the foreground.
In this example, the stereo information for the coplanar locations of the
dots does not prevent outline depiction, grouping, occlusion, and rever-
sals. Rather, it provides the conditions under which outline can operate.
The oblique bars, depicting corners, in the stereo-defined depicted cube
change orientation as the “+” intersections reverse depth and overlap.
354 John M. Kennedy, Igor Juricevic, and Juan Bai

Evidently the referents of outline are available to stereo-defined lines. If


outline is a grouping function, as hypothesized, the same functions should
be available to pictures of cubes available via motion-defined lines.

14.15 CONCLUSION

Common features of lines and surface edges allow outline drawings to


stand for edges of surfaces. An outline drawing generally uses a continu-
ous, polarized border. But dots induce lines that have no polarization.
These function as outline, even in stereo-defined lines, and, by hypothesis,
in motion-defined lines too.
The referents of outline are edges of surfaces. These are also defined
optically by continuous and discontinuous borders, and have no polariza-
tion. In sum, grouping effects in perception link physically separate ele-
ments of texture. These grouping effects serve surface perception and
outline.
Chapter 15
Irreconcilable Views

Hermann Kalkofen

Multiple views in perspectival pictures have been considered both fash-


ionable and irreconcilable. The history of such views deserves to be remem-
bered. Most of the contradictive views have arisen from violations of the
tacit “one-point-in-time-one-center-of-projection” canon of linear perspec-
tive depiction. Simply stated, this canon is characterized by three main
ingredients: “The three principal elements of a perspective construction
are,” in the words of Gezienus Ten Doesschate, “a the centre of perspec-
tive (the eye); b the picture plane; c the object” (Ten Doesschate 1964,
p. 24). This chapter deals with a and c; the terms “perspective” and “pro-
jection” will—unless otherwise declared—be used synonymously. I
should beg pardon in advance for a now-and-then unmannerly treatment
of works of art.

15.1 “VISUAL FIELDS”

Figure 15.1 may not exactly be considered a work of art; it is borrowed


from an illustrated history of psychology and is of the physicist, philoso-
pher, and psychologist Ernst Mach in his study in Prague, probably in the
1880s (Mach 1903). An ordinary photograph, a faithful picture, “erschei-
nungstreues Bild vom Typ der Photographie,” as Karl Bühler (1925,
p. 113) put it. Photography implies one point of view, one moment. A
drawing is a handmade picture, a chirograph in the sense of James J.
Gibson (1950), not inevitably faithful.
Figure 15.2 also depicts Mach in his study, about 1870. He was at
that time concerned with his phenomenal ego, the Selbstanschauung, in
the case of monocular vision. The outcome is this rather humorously
meant self-portrait “from within.” It was published by the draftsman in
356 Hermann Kalkofen

Figure 15.1
Ernst Mach. From J. Brozek and J. Hoskovec (1993, p. 229).

1885 in his Analysis of the Sensations and it has been revived in Gibson’s
Perception of the Visual World (1950). Gibson explained the portrayal
as a “literal representation” of Mach’s “visual field, with his right eye
closed, as he reclined in a nineteenth century chaise longue. His nose de-
limits the field on the right and his mustache appears below. His body
and the room are drawn in detail, although he could not see them in
detail without moving his eye” (Gibson 1950, p. 27). Mach’s portrait
from within represents what Hermann von Helmholtz called a Blickfeld,
whereby the eye is allowed here to move in its socket. Figure 15.3
shows the visual field of an arrested eye—in Helmholtz’s terms a Sehfeld
(see Metzger 1975, p. 127)—on top of Mach’s Blickfeld. The boundary
lines indicate appearance change as a function of where in the visual
field an object is placed. For example, a yellowish-green perimetric target
357 Irreconcilable Views

Figure 15.2
Self-portrait “from within.” From Mach (1903, p. 15).

appears, when first detected, whitish (solid line) and thereafter yellow
(broken line), and not until it reaches the dot-surrounded region does it
appear yellowish green. The so-called functional visual field is defined as
the area wherein some pattern discrimination with an arrested eye is just
possible. It has a horizontal extension of about 15° (Ikeda and Takeuchi
1975).1

15.2 CORNELIUS ON CLOSE DISTANCE CONSTRUCTIONS

Mach’s portrait from within comprises a generous viewing angle of 150°


horizontally. It is an extremely close-distance construction. Close distance
constructions of pictures have been firmly condemned as a kind of pro-
fessional blunder by Hans Cornelius, philosopher, psychologist, as well as
358 Hermann Kalkofen

Figure 15.3
A Sehfeld on top of Mach’s Blickfeld. From Mach (1903, p. 15) and Metzger
(1975, p. 127).

art historian and educator. In his Elementargesetze der bildenden Kunst


(Elementary laws of the visual arts) (1908),2 such constructions came to
be banned as they prevent the spectator from achieving a unified view, a
Fernbild, in the sense of Adolf Hildebrand (1908). Instead they merely cre-
ate a sum of disjointed impressions; they suffer from the flaw of multiple
perspectives (mehrfache Perspektive, Cornelius 1908, pp. 23–24). For
example, according to William Mackavey a picture as shown in figure 15.4
could just not create “a unitary and stable pictorial conception,” because
it provides “not only the intended impression of looking at or toward the
subject but will contain downward and upward views as well” (Mackavey
1980, p. 215). This pinhole photograph by Maurice Pirenne (1970) of a
Greek statuette covers (looked at from the projection center) a visual
angle of 90°, as do the deterrent examples that Cornelius (1908, p. 25)
constructed. What, then, is the visual angle that would in the eyes of
authors such as Cornelius permit a unified view at one glance? Neither
Cornelius nor Mackavey explicitly declared a number of degrees as the
359 Irreconcilable Views

Figure 15.4
Pinhole photograph of a Greek statuette (probably Aphrodite, second century
BC, according to Pirenne). From Pirenne (1970, p. 101).
360 Hermann Kalkofen

limit. The value recently declared by Michael Bischoff (1998, pp. 143–
144), how-ever, amounts to 36°. Outside this visual angle, according
to him, objects are no longer clearly perceptible with an unmoving eye.
Bischoff traces the 36° back to Hans Jantzen (1911), who asked for the
limit up to which a picture can be clearly seen with an unmoving eye but
beyond which several shifts of gaze are required to make it perceptible
(ibid., p. 120). Jantzen’s question remains, strictly speaking, unanswered;
he suspected the region to ultimately extend inside the area beyond which
increasing marginal distortions emerge. This area has been defined by
Guido Hauck, to whom Jantzen referred. On the basis of his own per-
spective constructions, Hauck estimated it to be between 30° and, maxi-
mally, 36°.

15.3 HAUCK ON SUBJECTIVE PERSPECTIVE

Hauck was not actually concerned with gaze direction shifts. He stood out
as a discerning critic of the geometrical-perspective teachings of his days,
which were still based, as he saw it, on the physiology of the Kepplers (sic)
and [Christoph] Scheiners. The latter considered the eye to be a resting
camera obscura, and yet they were certain of a direct mental apprehen-
sion of the retinal image as an instantaneous whole (Hauck 1879, p. 4).
The fact that this teaching, in 1879, had remained absolutely untouched
by the recent achievements of physiological optics might have been, in
Hauck’s eyes, at least partly, explained by the accident that the alleged
supreme fidelity of perspective received apparent confirmation by the
camera obscura pictures of photography (ibid., p. 9).3 Seeing, however,
consists in the eye ceaselessly moving up and down, fixating, and flying
over the integral object. According to Hauck, the skillfulness of the eye
in this traveling activity is so great that we do not in the least realize the
particulars of the process. In a trice, the total image (Gesammtbild) is cre-
ated, which we believe to be a combination of simultaneous detail impres-
sions, whereas in reality the same take place in succession (ibid., pp. 7–8).
Seeing means a transference of the successive to simultaneity. The result-
ing mental visual image (Hauck’s Gesammt-Anschauungsbild) would be
composed of the received detail impressions, in part contradicting each
other, whereas the duty to weigh and compare the importance of the
distinct detail impressions fell upon the mind. This is necessary in order
to even out the inconsistencies and to bring about an overall result that is
361 Irreconcilable Views

Figure 15.5
Fourfold row of pillars. From Hauck (1879).

consistent with itself and with the mind (ibid., p. 38). This is Helmholtzian
“cognitivist” thinking at its best.
The mathematician Hauck considered it thinkable that central perspec-
tive might represent just one out of an entire order of imaginable and
justifiable perspective systems (ibid., p. 6). However, he argued, “If we . . .
ask about the particular properties of our mental visual image two cardi-
nal features turn out: 1) The apparent size of a length is proportional
to the angle of view.—2) Any straight line appears again as straight line.”
The former feature would be called the principle of conformity, the latter
the principle of collinearity (ibid., pp. 39–40).
Being the aesthetician that he was, Hauck suggested that a graphic
representation, in order to equal the mental visual image as faithfully as
possible, should be endowed with the following features: (1) two indis-
pensable conditions to be fulfilled at all cost—the principle of the linear
horizon and the principle of verticality; (2) two secondary conditions to
be fulfilled if possible: the principle of conformity and the principle of
collinearity.4 The latter two conditions oppose each other and cannot be
satisfied at the same time; the one excludes the other completely or in part
(ibid., pp. 41–42).
Figure 15.5 shows a 95° view of a fourfold row of eight equidistant pil-
lars, whose appearance follows the rules of conform perspective and is
reshaped according to the verticality principle.5 Mind the perceptible cur-
vatures at top and bottom.
362 Hermann Kalkofen

Figure 15.6
Varieties of perspective. From Hauck (1879).

15.4 HAUCK ON ABSOLUTE PERSPECTIVE

Figure 15.6 compares the twelve pillars at the left in figure 15.5 delin-
eated in revised conform perspective (above) and in collinear perspective
(below). In both cases, the visual angle is 36°. Up to this angle, the dif-
ferences between conform and collinear pictures are so minute that it
might be considered a region of absolute perspective. The deviations would
further diminish if the angle becomes 30°. An angle of 30° corresponds,
however, to an eye distance just 2 times the width of the picture, an angle
of 36° to a distance of 1.5 picture widths,6 distances generally used and
recommended (ibid., p. 76). Is Hauck’s estimate of the absolute perspec-
363 Irreconcilable Views

Figure 15.7
Hyperbolic checkerboard pattern. From Helmholtz (1910).

tive area, being based on a comparison of a linear perspective picture with


a (no more than in its horizontal part) curvilinear perspective picture, an
overestimation?
Figure 15.7 shows the so-called curvilinear checkerboard pattern, a well-
known drawing by Helmholtz that is founded on a stereographic projec-
tion of the longitudinal and latitudinal meridians of the visual field, in the
sense of Blickfeld.7 The A is the appropriate eye distance (try to view the
figure from that distance!) that corresponds to a visual angle of 90° or
slightly more. The r (added by me) indicates the radius of that concentric
circle that comprehends, again, Hauck’s 36°. The answer to the question
of whether linearity is upheld inside the region thereby enclosed may be
a matter of taste; to me it tastes not too bad. So much for Hauck’s esti-
mate of the area in which artificial and natural perspective may peacefully
coexist.

15.5 PIRENNE ON EYE ROTATION AROUND A PROJECTION CENTER

The Jantzen’s question, however, needs further elaboration: Hauck’s 36°


area is still considerably larger than that of the functional visual field.
Even a 30° projection cannot be grasped with one glimpse; “distinct detail
364 Hermann Kalkofen

impressions” have to be successively collected and integrated into a men-


tal visual image here, too. Doing so clearly requires the apprehension of
the vast majority of perspectival pictures (when looked at properly from
their projection centers), that is, multiple views—but are they irreconcil-
able views? When dealing with this problem, Pirenne thinks of a certain
window, commonly associated with the name of Leonardo da Vinci but
probably invented by Alberti:
“Shutting one eye and keeping my head steady, I can, in principle at least, draw
on my study window a picture, in exact linear perspective, of the view outside
my window, by tracing (say, with a grease pencil) the outlines of all the objects
I see on the glass pane of the window.—Note that, keeping my head steady,
I could do this on different ‘windows,’ placed in different positions, or on win-
dows whose shape is curved instead of plane. Yet, all these outlines in perspective
would be seen to cover each other and to cover exactly the real objects, by virtue
of their very construction. . . . This is so even though the eye is continually rotat-
ing around a fixed center of rotation, because the images projected successively
on the retina, as the eye is rotating, always remain superimposed on one another
for each position of the eye. Contrary to widespread misconceptions, the physical
and physiological basis of linear perspective does not in the least entail the immo-
bility of the eye” (Pirenne 1975, pp. 438– 439; emphasis added). Let us, after
Pirenne’s authoritative plea, conclude this paragraph: even disjointed views may
be considered reconcilable in the case that they originate in a selfsame projection
center.

15.6 ZOOGRAPHY, SCENOGRAPHY

The third of Ten Doesschate’s principal elements of perspectival construc-


tion, c, was the object, plainly “the object.” Concerning narrative objects,
Seymour Chatman recognizes in his book Story and Discourse first of all,
a what and a how. “The story is the what in a narrative that is depicted,8
discourse the how” (Chatman 1978, p. 19). Within the story Chatman
discerns events and existents. Events comprise actions and happenings;
existents are made up of characters and the setting.9 In a more epistemo-
logical fashion, John Lyons (1977, p. 483) distinguishes between static
and dynamic situations.10 A static situation is one that is conceived of as
existing, rather than happening, and as being homogeneous, continuous,
and unchanging throughout its duration. A dynamic situation, on the
other hand, is something that happens or occurs. It might be momen-
tary or enduring, and it might or might not be under the control of an
agent.11 “If a dynamic situation is extended in time, it is a process . . . if
365 Irreconcilable Views

it is momentary, it is an event . . . and, if it is under the control of an


agent, it is an action. . . . Finally, a process that is under the control
of an agent is an activity; and an event that is under the control of an
agent is an act” (ibid., p. 483; asterisks omitted). Lyons’s static situations
equal Chatman’s existents; Chatman’s events resemble Lyons’s dynamic
situations.
Chatman’s discernment of characters and setting within the class of
existents brings us back to the notion of two originally separated artis-
tic disciplines: zoography, which is accountable for the depiction of living
beings, the characters, as Chatman puts it (or, the figures); and then scenog-
raphy, which is obliged to the depiction of the scene, the setting, to use
Chatman’s term (or, the ground). The two disciplines appear connected
first in Greece at the end of the third-century B.C.12 The double-purpose
utilization of pictorial space had begun. Characters (or, respectively,
figures) and setting (respectively ground) were nonetheless disparately
treated in Renaissance perspective painting, as has been pointed out by
Pirenne (1970, p. 121 and n. 6). Margaret Hagen comments on the same
issue: “Severe diminution occasionally may be tolerated across a scene,
but almost never across an object, particularly not across the human
figure” (Hagen 1986, pp. 136–137). This disparity is uniquely exemplified
in Andrea Mantegna’s impressive Mourning of Dead Christ (between 1485
and 1495). The photographic analysis of a secular model of the depicted
scene has been carried out by Robert Smith (1974), who shows that
“although the perspective diminution across the slab on which Christ’s
body rests is appropriate to a very close station point, approximately one-
and-a-half meters, the figure of Christ is not similarly foreshortened. The
proportions of the figure are consonant with a much longer [sic] station
point” (Hagen 1986, p. 137). The viewing distance comes, according to
Smith’s diagram (1974, p. 241), to about 21 m. The zoographic view of the
corpus Christi and the scenographic view of the slab are—in the prosaic
sense of this chapter—irreconcilable views.
Figure 15.8 is a chronophotograph of Etienne Jules Marey (1882;
cited after Ceram 1965, fig. 170). One is tempted to say that it reveals
the ontological validity of Chatman’s and Lyons’s discernments. The
chronophotograph was taken at ten equidistant points in time, t1–t10 . We
see one setting, the jumping posts and the bar, extending from t1–t10 , and
ten times the agent. The saying that perspective representation reduces the
three dimensions of space to two and the dimension of time to a point
366 Hermann Kalkofen

Figure 15.8
Chronophotograph of a high jump. From Ceram (1965, pp. 126–129).

is—considering time—apparently not equally valid for characters and


settings. Recording at multiple points in time, intentionally capturing
more than one moment, is not a chrono-photographic specificity. We shall
revisit this point in section 15.9.

15.7 PICTOGRAPHY, IDEOGRAPHY

What then is a faithful picture? Bühler’s definition of “erscheinungstreues


Bild,” translated here, faut de mieux, as “faithful,” is met in the photo-
graph as well as in “any picture constructed in accordance with the prin-
ciples of photographic representation, that means first and foremost,
which takes perspective (‘Formen-und Größenperspektive’), the distribu-
tion of light and shadow etc, into account” (Bühler 1925, p. 113). This
definition is possibly not in line with the thinking of Hauck (1879, p. 9);
quite noteworthy is, however, Bühler’s continuation: “whether the object
depicted really exists or not, whether the artist draws it on the basis of a
perception or on the basis of an idea, whether he reproduces it, as he had
367 Irreconcilable Views

Figure 15.9
Prisms and double prism. From Thiéry (1895, p. 319).

seen it, or whether recreating imagination had been in the game—all that
is insubstantial regarding the conception of pictorial faithfulness” (Bühler
1925, p. 113). Depictions of really existing (referent) objects are, semi-
otically spoken, pictographs; pictures of imaginary (referent) objects are
ideographs.

15.8 GOMBRICH ON THIÉRY’S FIGURE

An unsettling example of a picture containing multiple views was devised


by Armand Thiéry (1895) in his chirographic illustration shown in figure
15.9C (ibid., fig. 6). It is known as Thiéry’s figure. It was probably not
intended as a work of art but is nonetheless referred to in Gombrich’s Art
and Illusion: “It is practically impossible to keep this figure fixed because
it presents contradictory cues. The result is that the frequent reversals
force our attention to the plane. Thiéry’s figure, I believe, presents the
quintessence of cubism” (Gombrich 1983, p. 241). The name Armand
Thiéry is pair-associated with the perspective theory of the geometri-
cal size illusions. The theory posits that the illusion comes about when
perspective suggests depth and thereby effects size changes (Gregory
1977, p. 150). Thiéry’s Pseudoskopisches Doppelprisma (Wundt’s term),
is an improvement, as it were, of the Zöllner illusion, which Thiéry had
attempted to explain. It is conceived as composed of drawings of two
368 Hermann Kalkofen

Figure 15.10
From Theodor De Bry’s Collectiones peregrinatorum in Indiam Orientalem et
Indiam Occidentalem (1590–1634) (1990, p. 27). See plate 5 for color version.

prisms (figures 15.9A and 15.9B; figs. 4 and 5 in Thiéry 1895);13 the one at
the left seen from below, the right one seen from above. If both get com-
bined and consequently have a lateral plane in common, an ambiguous
pattern results, which can be interpreted in a twofold way: “If one moves
one’s eyes from the right to the left, the plane in common with its right
edge comes to the fore and forms, with the plane at the right, the prism
seen from above. The same right edge recedes, however, when the eyes run
through the figure from the second to the third plane to the left. In this
case, the planes located at the left apparently portray the prism seen from
below” (ibid., p. 320). The two interpretations of Thiéry’s figure go along
with irreconcilable views.14

15.9 MULTIPLE MOMENTS

Prisms and double-prisms, too, are a variety of inanimate object, some sort
of existent. The depiction of an existent may be called its presentation.
369 Irreconcilable Views

Figure 15.11
Noble Maiden from Segota. From De Bry (1990, p. 23).

Figure 15.10 (see also plate 5) shows an engraving from Theodor De


Bry’s Collectiones peregrinatorum in Indiam Orientalem et Indiam Occiden-
talem (1590–1634; see De Bry 1990). Is it the picture of two views of the
same woman, or are there two women presented? The German title of the
engraving may be translated as How the Women of Dasamonquepeuc Are
Accustomed to Carrying Their Children. One is tempted to take it as an
instructive illustration presenting the front and back of the same person.
However, the differential coloring of skin and cloth casts doubt on the
unity of the character.15
Figure 15.11 brings another example from De Bry’s Collectiones. The
two figures are differently colored, too, but according to the picture’s title
they depict one Noble Maiden from Segota; both are actually meant as
presentations of the same character. The two views of the girl can be
identified as one point-of-view (POV) shot, which, in this case, shows the
canonical form of the object, and as one reverse angle (RVA) shot, which
shows the rear view (Salt 1983). This opposition applies, however, to the
character only and not to the setting, with its zoographic interspersions
370 Hermann Kalkofen

Figure 15.12
Two Venetian Women, Albrecht Dürer (1495). From Strieder (1996, p. 150).

that are both contained in the POV shot.16 This construction of the—then
composite—picture requires two centers of projection—one for a zoo-
graphic view and one for a “holistic” view. The moment intended here
corresponds to one point in time. The alternative construction assumes
a single center of projection and two points in time.17 Here the noble
maiden had to change her statuary position during the time lapse between
t1 and t2 exactly into its opposite.
Figure 15.12 is a drawing by Albrecht Dürer (1495). It is admittedly
true that its caption reads as Two Venetian Women. The picture shall be
nonetheless regarded as presentation of a single character. The sceno-
371 Irreconcilable Views

Figure 15.13
Angela Merkel. From Die Zeit, no. 19, 2000. Photographs by Uta Mahler with
permission by Ostkreuz Agentur der Fotografen.

Figure 15.14
From De Bry (1990, p. 156).
372 Hermann Kalkofen

Figure 15.15
From Stemberger (1990).

graphic cues denoting a unified pictorial space are minimally elaborated.


The zoograph at the left presents the character in its canonical form; the
one at the right shows its rear view. How is the distance in depth between
the two views to be interpreted? Does it result from locomotion, that is,
from an (ecliptically depicted) action performed by the character? And
if so, how did the action proceed? Does the canonic view designate its
beginning or rather the end? Either way, this construction assumes two
points in time and one projection center. The alternative assumes, again,
one point in time and two projection centers.
373 Irreconcilable Views

Figure 15.16
From Racinet (1888/1995, p. 85).

Figure 15.13 brings us a composite photograph presenting two zoo-


graphic views of a contemporary German politician. Any scenographic
cues have been obliterated here. Clearly, there are two points in time des-
ignated. The difference in projective size of the views could in principle
have been effected by locomotion of the object. More likely, however, it
was achieved by traveling of the projection center, and even more likely by
the surrogate manipulation of a zoom lens. The presentation comprises,
anyhow, two points of view, two points in time. The order of the moments
remains undecided.
Narration proper needs chronology. The heading of De Bry’s engrav-
ing presented in figure 15.14 reads: The Wife of a King in the Province of
Cumana Brings Gifts to the Landvogt Herrera (1594). The temporal order
of the two views suggested is from the right to the left. She approaches
from the middle ground at the right, t1, and has arrived in the foreground,
t2. Order in time and location in space are congruous here; remoteness in
space corresponds in this case with remoteness in time passed. The per-
spective may be named retrospective (Kalkofen 1993).
374 Hermann Kalkofen

Figure 15.17
From Kratzsch (1983, p. 39).

However, future time, too, is remote and distant from the present. The
picture Serveto and His Execution (Christof van Sichem, 1608) (figure
15.15) is an instance of what may be named prospective perspective (ibid.).
Figure 15.16 brings a detail from a Japanese woodcut The Fitting out
of an Archer. It is the selfsame soldier indeed in both views, as can be
deduced from the omitted context. The scenographic cues are minimized
again. Pictorial space is reduced to the foreground. Thus, the depicted
action proceeds from the left to the right.18 Figure 15.17 shows a woodcut
from the Luther Bible (Kratzsch 1983), Samson Beats the Philistines, 1534.
This is another case of prospective perspective.
Pictorial narration may of course extend to more than two bare
moments. For example, De Bry’s engraving presented in figure 15.18
(Indians want to try out whether the Spaniards be immortal folks and drown
a Spaniard named Salsedo in the sea) comprises three moments and is
another example of prospective perspective. The instances of Wickhoff’s
continuous style with its characteristic figure repetitions could be enor-
375 Irreconcilable Views

Figure 15.18
From De Bry (1990, p. 158).

mously augmented by further examples. Suffice it to say that the rule of


the continuous style lasted, as stated by Franz Wickhoff, through one and
a half millennia, until the sixteenth century (Wickhoff 1912, p. 13). The
disappearance of continuous style began a long time before the inven-
tion of photography proper. Multiple use of scenographic space for zoo-
graphic purposes became unfashionable—for whatsoever reasons, not
necessarily aesthetic ones. Frescoes worked out by Sandro Botticelli,
Domenico Ghiralandaio, and Luca Signorelli in the continuous manner
beautify through today the walls of the Sistine Chapel.

15.10 EGYPTIAN VIEWS

The so-called Egyptian method (Arnheim 1974, p. 112) may be considered


as another, primordial, attempt in line of the multiple use of pictorial
376 Hermann Kalkofen

Figure 15.19
From Lübke and Pernice (1921, fig. 71/51), Sethos I. Offers a Picture to Truth
(Seth, Seti), nineteenth dynasty (1314–1200 B.C.).

space, which is here, of course, at most zoographical space. The famous


drawing by Daniel Alain (Gombrich 1983, fig. 1; Gibson 1979, fig. 15.3) is
a perceptive cartoon, perceptive also in that it represents the model’s right
arm as ending in a left hand. A similar case has been made by Pirenne
(1970, p. 79) for an authentic papyrus drawing, where the left arm of the
figure ends in a “wrong” hand. Unbeatable in this regard, however, is this
presentation of Sethos I, (figure 15.19), the father of Ramses II. Let me
conclude: not only the abolition of the one-point-in-time-one-center-of-
projection canon leads to irreconcilable views.
377 Irreconcilable Views

Notes
1. The term has seemingly been coined by Sanders (1970); Ikeda and Takeuchi
report that “the functional visual field varied greatly among subjects, and also
changed for a particular subject when the data were collected on different occa-
sions” (1975, p. 260).
2. The author was among the founding fathers of Gestalt theory (cf. Metzger
1975, p. 133). Jantzen (1911, p. 120) characterized Cornelius’ book as “reich illus-
trierten Kommentar zu Adolf Hildebrands Problem der Form” [“richly illustrated
commentary on Adolf Hildebrand’s Problem of Form”] (Hildebrand 1908).
3. Compare Te Peerdt: “Vor Erfindung der Photographie ist nicht einmal der
Gedanke einer wahrhaft einheitlichen optischen Darstellung gefasst gewesen”
[“Before the invention of photography the thought of a truly consistent optical
representation had not even existed”] (1899, p. 6).
4. I have reversed the order of the two principles from the order in which he gave
them.
5. Hauck’s “reconceived pictorial space”—one system of “curvilinear perspec-
tive” among others (Pirenne 1970, p. 148)—has been assessed by Ten Doesschate:
“He introduced a new system of perspective which he called ‘conform’ in opposi-
tion to normal central perspective which is collinear.—Hauck is not consistent.
He represents horizontal straight lines by arcs; verticals, however, he ‘rectifies.’
This he does according to our ‘consciousness of verticality’ (Vertikalitätsbewusst-
sein)” (1964, pp. 47– 48). Hauck’s semiconformity has already been criticized by
Panofsky (1927, p. 262).
6. In the case of still smaller viewing distances, it would be feasible to begin the
construction collinearily first and to reshape possible conformity distortions later
on, according to conform perspective. Raphael always proceeded in this way, see,
for example, the spheres at the right and the base of the column at the left in The
School of Athens, 1510–11, (Hauck 1879, p. 75). Compare Pirenne: “The architec-
ture which extends over most of the painting is drawn in perspective as one whole,
from one main centre of projection. But the spheres (and the numerous human
figures) are not drawn as projections from this centre. They are drawn from a
number of subsidiary centres of projection, each in front of the position which the
respective sphere or figure would occupy in the painting” (1970, p. 121).
7. “[D]ie Richtkreise des Blickfeldes, welche mit der durch den Fixationspunkt
gehenden vertikalen und horizontalen Linie übereinstimmende Richtung haben,
auf eine ebene Tafel projiziert” [“the reference circles that have a common direc-
tion with the vertical and horizontal lines that cross the fixation point, projected
onto a plane”] (Helmholtz 1910, p. 150).
8. “[P]icture narratives have, of course, been common for centuries” (Chatman
1978, p. 34).
9. Chatman (1978) notes that this kind of distinction has been recognized since
antiquity, that is, in Aristotle’s Poetics (p. 19). Comparably, G. E. Lessing drew a
distinction between characters (Körper) and actions (Handlungen) (Lessing 1964).
378 Hermann Kalkofen

10. “The conceptual framework within which we organize and describe our per-
ceptions of the physical world, whatever language we speak, is one in which we can
identify not only states-of-affairs of shorter or longer duration, but also events,
processes and actions.—There is, unfortunately, no satisfactory term that would
cover states, on the one hand, and events, processes and actions, on the other”
(Lyons 1977, p. 483). Lyons’ opposition reminds us of L. von Bertalanffy’s (1952)
saying that “what are called structures are slow processes.” Consistent with this
opposition is Strawson’s ontological distinction between first- and second-order
entities (Lyons 1977, p. 443).
11. An agent is, in the sense of Chatman, an existent; in Lyons’s sense he is a
static situation. Lyons refrains—understandably—from naming agents situations.
12. In the words of Franz Wickhoff: “Endlich, für unsere gewohnte Vorstel-
lung von der Kunstentwicklung spät genug, hatte sich auch im Bilde zusam-
mengeschlossen, was man auf der Szene schon lange zu sehen gewohnt war:
wandelnde Figuren vor einem geschlossenen Grund” [“Late enough for our accus-
tomed idea of the development of art, something long familiar on the scene finally
came together for the picture: changing figures in front of a closed ground”] (1912,
p. 104).
13. This is a somewhat strange composition in that Thiéry’s figures 4, 5, and 6
show nine, eight, and ten crossing transversals respectively.
14. The picture-prisms are, of course, perspective projections: Thiéry’s drawing
instructions for the prisms read as follows: “Man denke sich ein verticales Prisma,
von dem bloß zwei Seitenflächen gesehen werden. Das Prisma ist um eine hori-
zontale Achse beweglich. Auf jede der sichtbaren Flächen des Prismas zeichnet
man eine verticale durch horizontale Striche eingetheilte Linie. Wenn das Prisma
so gedreht worden ist, dass sein oberer Theil vom Beobachter entfernt liegt, dann
wird das Prisma als Figur 4 und die wagerechten Striche werden als Transversalen
erscheinen. Die eingetheilten Linien werden jetzt in Folge dessen nach oben con-
vergent gesehen. Da das Auge des Beobachters in der Höhe der Achse liegt, so sind
auch die oberen Eckpunkte der Linien dem Beobachter näher als die unteren;
der Gesichtswinkel der Distanz zwischen den oberen Endpunkten ist also kleiner
wie der zwischen den unteren. Ferner: wirklich nach oben divergirende Linien
würden in diesem Fall als parallele Linien gezeichnet werden müssen” (Thiéry
1895, pp. 321–322).
15. The differential coloring is probably an indication of the advanced decay of
what has been called by Wickhoff the continuous style, with its characteristic figure
repetitions.
16. Or in the RVA shot: the picture can be “read” alternately.
17. Compare Vitz’s nice analysis of Edouard Manet’s Bar at the Folies Bergère,
1881 (Vitz 1979, p. 139).
18. The action proceeds with varying time intervals. Its depiction relates, in con-
trast to Marey’s chronophotography of the jump, to an intrinsic segmentation.
References

Adams, P. A., and Haire, M. (1959). The effect of orientation on the reversal of
one cube inscribed in another. American Journal of Psychology, 72, 296–299.
Adelson, E. H. (1993). Perceptual organization and the judgement of brightness.
Science, 262, 2042–2044.
Alberti, L. B. (1436/1966). On painting (trans. J. R. Spencer). New Haven, Conn.:
Yale University Press.
Alberti, L. B. (1436/1972). On painting (trans. C. Grayson). London: Phaidon.
Alhazen, I. (1039/1989). Book of optics. In A. I. Sabra (ed.), The optics of Ibn
Al-Haytham. London: Warburg Institute, University of London. (Alhazen died
about 1039 C. E. When he wrote this work is not known.)
Anderson, C. H., and Van Essen, D. C. (1987). Shifter circuits: A computational
strategy for dynamic aspects of visual processing. Proceedings of the National
Academy of Science, 84, 6297–6301.
Anderson, K. (1992). Brook Taylor’s work on linear perspective. Sources in the
history of mathematics and physical sciences, vol. 10. New York: Springer.
Arend, L., and Goldstein, R. (1987). Simultaneous constancy, lightness and bright-
ness. Journal of the Optical Society of America A, 4, 2281–2285.
Arnheim, R. (1974). Art and visual perception: A psychology of the creative eye.
Berkeley: University of California Press.
Atkins, J. E., Fiser, J., and Jacobs, R. A. (2001). Experience-dependent visual
cue integration based on consistencies between visual and haptic percepts. Vision
Research, 41, 449–461.
Attneave, F. (1954). Some informational aspects of visual perception. Psycholog-
ical Review, 61, 183–193.
Bahn, P., and Vertut, J. (1997). Journey through the Ice Age. Berkeley: University
of California Press.
Ballard, D. (1991). Animate vision, Artificial Intelligence, 48, 57–86.
Baxandall, M. (1985). Patterns of intention: On the historical explanation of pic-
tures. London: Yale University Press.
380 References

Bazin, A. (1967). The ontology of the photographic image. In What is cinema?,


vol. 1, selected and translated by Hugh Gray. Berkeley and Los Angeles: Univer-
sity of California Press.
Bechtel, W., and Abrahamsen, A. (1991). Connectionism and the mind: An intro-
duction to parallel processing in networks. Cambridge, MA: Blackwell.
Belhumeur, P. N. (1996). A computational theory for binocular stereopsis. In
D. C. Knill and W. Richards (eds.), Perception as Bayesian inference (pp. 323–
364). New York: Cambridge University Press.
Belhumeur, P., Kriegman, D. J., and Yuille, A. L. (1999). The bas-relief ambigu-
ity. International Journal of Computer Vision, 35, 33–44.
Berger, J. (1972). Ways of seeing. London: BBC and Penguin.
Berkeley, G. (1709/1975). An Essay toward a new theory of vision. In Philosoph-
ical works, introduction and notes by M. R. Ayers. London: J. M. Dent and Sons.
(Original work appeared in 1709, Dublin.)
Bertalanffy, L. von (1952). Problems of life: An evaluation of modern biological
thought. London: Watts & Co.
Biederman, I. (1985) Human image understanding: Recent research and a theory.
Computer Vision, Graphics and Image processing, 32, 29–73.
Birnbaum, M. H. (1983). Scale convergence as a principle for the study of percep-
tion. In H. Geissler (ed.), Modern issues in perception (pp. 319–335). Amsterdam:
North-Holland.
Bischoff, M. (1998). Nahdistanzkonstruktion und Bildwahrnehmung (Construc-
tion and Picture Perception). In K. Sachs-Hombach and K. Rehkämper (eds.),
Bild—Bildwahrnehmung—Bildverarbeitung (image—picture perception—image
processing) (pp. 143–151), Wiesbaden: Deutscher Universitätsverlag.
Bool, F. H., Kist, J. R., Locher, J. L., and Wierda, F. (1981). M. C. Escher: His
life and complete work. New York: Harry Abrams.
Born, M., and Wolf, E. (1959). Principles of optics. London: Pergamon.
Bouret, J. (1950). Picasso, dessins. Paris: Diffusions Française.
Boyd, R. (1979). Metaphor and theory change: What is “metaphor” a metaphor
for? In A. Ortony (ed.), Metaphor and thought (pp. 356–408). Cambridge: Cam-
bridge University Press.
Brentano, F. von (1874). Die Psychologie vom empirischen Standpunkte (Psychol-
ogy from an empirical standpoint). Leipzig, Germany: Duncker und Humbolt.
Brodatz, P. (1966). Textures. New York: Dover.
Brozek, J., and Hoskovec, J. (1993). Zur Geschichte der Prager deutschsprachi-
gen Psychologie (Regarding the History of the German Speaking Psychology in
Prague). In H. Lück and R. Miller (eds.), Illustrierte Geschichte der Psychologie
(Illustrated History of Psychology) (pp. 227–232). Munich: Quintessenz.
Bruno, N., and Cutting, J. E. (1988). Minimodularity and the perception of lay-
out. Journal of Experimental Psychology: General, 117, 161–170.
381 References

Buckley, D., Frisby, J. P., and Freeman, J. (1994). Lightness perception can be
affected by surface curvature from stereopsis. Perception, 23, 869–881.
Bühler, K. (1922). Die Erscheinungsweisen der Farben (The appearance of the
colors). In K. Bühler (ed.), Handbuch der Psychologie, part 1: Die Struktur der
Wahrnehmungen (Handbook of Psychology, part 1: The structure of Perception)
(pp. 1–201). Jena, Germany: Fischer.
Bühler, K. (1925). Abriß der geistigen Entwicklung des Kindes (Mental Develop-
ment of the Child). Leipzig, Germany: Quelle and Meyer.
Bunge, W. (1966). Theoretical geography. Lund, Sweden: C. W. K. Gleerup.
Burton, H. E. (1945). The optics of Euclid. Journal of the Optical Society of
America, 35, 357–372.
Busey, T. A., Brady, N. P., and Cutting, J. E. (1990). Compensation is unneces-
sary for the perception of faces in slanted pictures. Perception and Psychophysics,
48, 1–11.
Cabe, P. A., Abele, P. N., and Kennedy, J. M. (1999). Dominance of bino-
cular depth information over competing kinetic occlusion information. Paper
presented at the meeting of the Psychonomic Society, November 18–21, Los
Angeles.
Cadiou, H., and Gilou (1989). La peinture en trompe l’oeil (Trompe l’oeil Paint-
ing). Paris: Dessain et Tolra.
Carani, M. (1998). The semiotics of perspective, Semiotica, 119 (3/4), 309–357.
Carlbom, I., and Paciorek, J. (1978). Planar geometric projections and viewing
transformations. Computing Surveys, 10, 465–502.
Carl Zeiss Jena (1907). Instrument zum beidäugigen Betrachten von Gemälden
u.ggl (Instrument for the Binocular Viewing of Paintings). Kaiserliches Patentamt
(patent no), Patentschrift no. 194480, Klasse 42h, Gruppe 34.
Carroll, N. (1988). Mystifying movies: Fads and fallacies in contemporary film
theory. New York: Columbia University Press.
Cavanagh, P., and Kennedy, J. M. (2000). Close encounters: Small details veto
depth from shadows. Science, 287, 2421.
Cavanagh, P., and Leclerc, Y. G. (1989). Shape from shadows. Journal of Experi-
mental Psychology: Human Perception and Performance, 15, 3–27.
Cazeaux, C. (in press). Metaphor and the categorization of the senses. Metaphor
and Symbol.
Cavanagh, P. (1987). Reconstructing the third dimension: Interactions between
color, texture, motion, binocular disparity, and shape. Computer Vision, Graphics,
and Image Processing, 37, 171–195.
Ceram, C. W. (1965). Eine Archäologie des Kinos (An Archeology of Cinema).
Reinbek: Rowohlt.
Chatman, S. (1978). Story and discourse: Narrative structure in fiction and film.
Ithaca, N.Y.: Cornell University Press.
382 References

Chauvet, J.-M., Brunel Deschamps, E., and Hillaire, C. (1995). La Grotte Chauvet
à Vallon-Pont-d’Arc. Paris: Seuil.
Chomsky, N. (1957/1971). Syntactic structures. The Hague, Netherlands:
Mouton.
Chomsky, N. (1995). The minimalist program. Cambridge, Mass.: MIT Press.
Chomsky, N. (1996). Powers and prospects: Reflections on human nature and the
social order. London: Pluto Press
Chomsky, N. (2000). New horizons in the study of language and mind. Cambridge:
Cambridge University Press.
Churchland, P. M. (1989). Perceptual plasticity and theoretical neutrality: A reply
to Jerry Fodor, Philosophy of Science, 55, 167–187.
Churchland, P. M. (1992). A neurocomputational perspective: The nature of mind
and the structure of science. Cambridge, MA: MIT Press.
Churchland, P. S., Ramachandran, V. S., and Sejnowski, T. J. (1994). A critique
of pure vision. In C. Koch and J. C. Davis (eds.), Large-scale neuronal theories of
the brain (pp. 23–60). Cambridge, Mass.: MIT Press.
Clottes, J. (2001). La Grotte Chauvet: L’art des origines (The Art of the Origins).
Paris: Seuil.
Clowes, M. B. (1971). On seeing things. Artificial Intelligence, 2 (1), 79–116.
Cook, M. (1978). Judgment of distance on a plane surface. Perception and Psy-
chophysics, 23, 85–90.
Cornelius, H. (1908). Elementargesetze der bildenden Kunst: Grundlagen einer prak-
tischen Ästhetik (Elementary Laws of Visual Art: Foundations of a Practical
Aesthetics). Leipzig and Berlin: Teuber.
Costall, A. P. (1990). Seeing through pictures. Word and Image, 6, 273–277.
Cutting, J. E. (1986a). Perception with an eye for motion. Cambridge, Mass.: MIT
Press.
Cutting, J. E. (1986b). The shape and psychophysics of cinematic space. Behavior
Research Methods, Instruments, and Computers, 18, 551–558.
Cutting, J. E. (1987). Rigidity in cinema seen from the front row, side aisle. Journal
of Experimental Psychology: Human Perception and Performance, 13, 323–334.
Cutting, J. E. (1988). Affine distortions in pictorial space: Some predictions for
Goldstein (1987) that La Gournerie (1859) might have made. Journal of Experi-
mental Psychology: Human Perception and Performance, 14, 305–311.
Cutting, J. E. (1997). How the eye measures reality and virtual reality. Behavior
Research Methods, Instruments, and Computers, 29, 27–36.
Cutting, J. E. (1998). Information from the world around us. In J. Hochberg (ed.),
Perception and cognition at century’s end (pp. 69–93). New York: Academic Press.
Cutting, J. E., and Massironi, M. (1998). Pictures and their special status in per-
ceptual and cognitive inquiry. In J. Hochberg (ed.), Perception and cognition at
century’s end: History, philosophy, and theory (pp. 137–168). San Diego: Academic
Press.
383 References

Cutting, J. E., and Vishton, P. M. (1995). Perceiving layout and knowing dis-
tances: The integration, relative potency, and contextual use of different informa-
tion about depth. In W. Epstein and S. Rogers (eds.), Perception of space and
motion (pp. 69–117). San Diego, Calif.: Academic Press.
Damisch, H. (1994). The origin of perspective (trans. John Goodman). Cambridge,
Mass.: MIT Press.
Danto, A. (2001). Seeing and showing. Journal of Aesthetics and Art Criticism, 59,
1–10.
Da Silva, J. A. (1985). Scales of perceived egocentric distance in a large open field:
Comparison of three psychophysical methods. American Journal of Psychology,
98, 119–144.
De Bry, T. (1990). Reisen in das östliche und westliche Indien: Reisen in das west-
liche Indien (Amerika) (Travels to east and west Indies; Travels to the West Indies
(America)), ed. G. Sievenich. Berlin and New York: Casablanca Verlag.
Denis, M. (1890/1976). In R. Goldwater and M. Treves (eds.) Artists on art: From
the 14th to the 20th century (p. 380). London: John Murray.
Deregowski, J. B., Muldrow, E. S., and Muldrow, W. F. (1972). Pictorial recogni-
tion in a remote Ethiopian population. Perception, 1, 417– 425.
Descartes, R. (1637/1965). Discourse on method, optics, geometry, and meteorology
(trans. Paul J. Olscamp). New York: Bobbs-Merrill.
Diels, H. (1922). Die Fragmente der Vorsokratiker, 4th ed., vol. 1. Berlin: Weid-
mannsche Buchhandlung.
Doorn, A. J. van, Koenderink, J. J., Ridder, H. de (2001). Pictorial space corre-
spondence in photographs of an object in different poses. In B. E. Rogowitz and
T. N. Pappas (eds.), Proceedings of SPIE: Human vision and electronic imaging VI.
4299, 321–329.
Dosher, B. A., Sperling, G., and Wurst, S. A. (1986). Tradeoffs between stereop-
sis and proximity luminance covariance as determinants of perceived 3D structure,
Vision Research, 26, 973–990.
Dubery, F., and Willats, J. (1972). Drawing systems. London: Studio Vista.
Dubery, F., and Willats, J. (1983). Perspective and other drawing systems. London:
Herbert Press; New York: Van Nostrand Reinhold.
Eby, D. W., and Braunstein, M. L. (1995). The perceptual flattening of three-
dimensional scenes enclosed by a frame. Perception, 24, 981–993.
Epstein, W. (1968). Modification of the disparity-depth relationship as a result
of exposure to conflicting cues. The American Journal of Psychology, 81, 189–197.
Epstein, W. (1982). Percept-percept coupling. Perception, 11, 75–83.
Epstein, W., and Rogers, S. (in press). Percept-percept coupling revisited. In
U. Savardi (ed.), Festschrift in onore di Paolo Bozzi. Padua, Italy: Cleup.
Erens, R. G. F., Kappers, A. M. L., and Koenderink, J. J. (1991). Limits on the per-
ception of local shape from shading. In P. J. Beek, R. J. Bootsma, and P. C. W. van
Wieringen (eds.), Studies in perception and action (pp. 65–71). Amsterdam: Rodopi.
384 References

Erens, R. G. F., Kappers, A. M. L., and Koenderink, J. J. (1993). Perception of


local shape from shading. Perception and Psychophysics, 54, 145–156.
Evans, G. (1982). The varieties of reference, ed. J. McDowell. Oxford: Oxford
University Press.
Farber, J., and Rosinski, R. R. (1978). Geometrical transformations of pictured
space. Perception, 7, 269–282.
Faugeras, O. D. (1993). Three dimensional computer vision: A geometric viewpoint.
Cambridge, Mass.: MIT Press.
Faul, F. (1997). Theoretische und experimentelle Untersuchungen chromatischer
Determinanten perzeptueller Transparenz (Theoretical and experimental investiga-
tions of chromatic determinants of perceptual transparency). Dissertation thesis,
University of Kiel.
Field, J. V. (1993). Mathematics and the craft of painting: Piero della Francesca
and perspective. In J. V. Field and F. A. J. L. James (eds.), Renaissance and revo-
lution (pp. 73–95). Cambridge: Cambridge University Press.
Flavell, J. H., Flavell, E. R., and Green, F. L. (1983). Development of the
appearance-reality distinction. Cognitive Psychology, 15, 95–120.
Fleming, W. (1986). Arts and ideas. New York: Holt, Rinehart, and Winston.
Flückiger, M. (1991). La perception d’objets lointains. In M. Flückiger and
K. Klaue (eds.), La perception de l’environnement (pp. 221–238). Lausanne,
Switzerland: Delachaux et Niestlè.
Fodor, J. (1983). The modularity of mind. Cambridge, Mass.: MIT Press.
Freedberg, D. (1989). The power of images. Chicago: University of Chicago Press.
Fuchs, W. (1923a). Experimentelle Untersuchungen über das simultane Hinterein-
andersehen auf derselben Sehrichtung. Zeitschrift für Psychologie, 91, 145–235.
Fuchs, W. (1923b). Experimentelle Untersuchungen über die Änderung von Far-
ben unter dem Einfluß von Gestalten (“Angleichungserscheinungen”) (Experi-
mental investigations about the influence of Gestalts on the change of color).
Zeitschrift für Psychologie, 92, 249–325.
Gablik, S. (1985). Magritte. London: Thames and Hudson.
Galanter, E., and Galanter, P. (1973). Range estimates of distance visual stimuli.
Perception and Psychophysics, 14, 301–306.
Gallistel, C. R. (1998). Symbolic processes in the brain: The case of insect naviga-
tion. In D. Scarborough and S. Sternberg (eds.), Methods, models and conceptual
issues: An invitation to cognitive science, vol. 4 (pp. 1–51). Cambridge, Mass.: MIT
Press.
Gallistel, C. R. (2000). The replacement of general-purpose learning models with
adaptively specialized learning modules. In M. S. Gazzaniga (ed.), The cognitive
neurosciences, 2d. ed. (pp. 1179–1191). Cambridge, Mass.: MIT Press.
Gårling, T. (1970). Studies in visual perception of architectural spaces and rooms,
III: A relation of judged depth to judge size of space. Scandinavian Journal of
Psychology, 11, 124–131.
385 References

Gelb, A. (1921). Grundfragen der Wahrnehmungspsychologie (Basic questions in


perceptional psychology). In K. Bühler (ed.), Bericht über den VII. Kongreß für
experimentelle Psychologie (Report on the 7th congress of experimental psychol-
ogy) (pp. 114–115). Jena, Germany: Gustav Fischer.
Gelb, A. (1929). Die “Farbenkonstanz” der Sehdinge. In A. Bethe, G. v. Bergmann,
G. Embden, and A. Ellinger (Eds.), Handbuch der normalen und pathologischen
Physiologie. Bd.12, 1.Hälfte. Receptionsorgane II. (pp. 594678). Berlin: Springer.
[Transl: The color constancy of visual objects. Handbook of normal and patho-
logical physiology, vol. 12: Organs of reception]
Gelb, A., and Granit, R. (1923). Die Bedeutung von “Figur” und “Grund” für die
Farbenschwelle (The significance of figure and ground for the color threshold).
Zeitschrift für Psychologie, 93, 83–118.
Gibson, J. J. (1947/1982). Pictures as substitutes for visual realities. In E. Reed and
R. Jones (eds.), Selected essays of James J. Gibson (pp. 231–240). Hillsdale, N.J.:
Erlbaum. (Original work appeared in 1947.)
Gibson, J. J. (1950). The perception of the visual world. Boston: Houghton Mifflin.
Gibson, J. J. (1954). The visual perception of objective motion and subjective
movement. Psychological Review, 61, 304–314.
Gibson, J. J. (1954). A theory of pictorial perception. Audio-visual Communications
Review, 1 (3). (Gibson does not give any page nos. and I have been unable to track
them down.)
Gibson, J. J. (1960). Pictures, perspective, and perception, Dædalus, 89 (winter
issue), 216–227.
Gibson, J. J. (1966). The senses considered as perceptual systems. Boston:
Houghton Mifflin.
Gibson, J. J. (1971). The information available in pictures. Leonardo, 4, 27–35.
Gibson, J. J. (1979). The ecological approach to visual perception. Boston:
Houghton Mifflin.
Gibson, E. J., and Bergman, R. (1954). The effect of training on absolute esti-
mation of distance over the ground. Journal of Experimental Psychology, 48,
473–482.
Gibson, J. J., and Cornsweet, J. (1952). The perceived slant of visual surfaces—
optical and geographical. Journal of Experimental Psychology, 44, 11–15.
Gilden, D. (1991). On the origins of dynamical awareness. Psychological Review,
98, 554–568.
Gilinsky, A. S. (1951). Perceived size and distance in visual space. Psychological
Review, 58, 460–482.
Gillam, B. J. (1968). Perception of slant when perspective and stereopsis conflict:
Experiments with aniseikonic lenses. Journal of Experimental Psychology, 78,
299–305.
Gillam, B. J. (1995). The perception of spatial layout from static optical informa-
tion. In W. Epstein and S. Rogers (eds.), Perception of space and motion. New
York: Academic Press.
386 References

Gillam, B., and Sedgwick, H. A. (1996). The interaction of stereopsis and per-
spective in the perception of depth. Perception, 25 (supplement), 70.
Gillam, B., Sedgwick, H. A., and Cook, M. (1993). The interaction of surfaces
with each other and with discrete objects in stereoscopic depth. Perception, 22
(supplement), 35.
Glennerster, A., and McKee, S. P. (1999). Bias and sensitivity of stereo judge-
ments in the presence of a slanted reference plane. Vision Research, 39, 3057–
3069.
Gogel, W. C. (1993). The analysis of perceived space. In S. Masin (ed.), Founda-
tions of perceptual theory (pp. 113–182). Amsterdam: North-Holland.
Gogh, V. van (1959). The complete letters of Vincent van Gogh: With reproductions
of all the drawings in the correspondence, 2d ed., vol. 1. Greenwich, Conn.: New
York Graphic Society.
Goldsmith, T. H. (1990). Optimization, constraint, and history in the evolution
of eyes. The Quarterly Review of Biology, 65, 281–322.
Gombrich, E. (1956). Art and illusion: A study in the psychology of pictorial rep-
resentation. Princeton, N.J.: Princeton University Press.
Gombrich, E. H. (1959). Art and illusion, part 3: The beholder’s share. London:
Phaidon Press.
Gombrich, E. H. (1960). Art and illusion: A study in the psychology of pictorial
representation. New York: Pantheon.
Gombrich, E. H. (1960/1993). Art and illusion: A study in the psychology of picto-
rial representation. London: Phaidon.
Gombrich, E. H. (1974). “The sky is the limit”: The vault of heaven and pictorial
vision. In E. H. Gombrich, The image and the eye: Further studies in the psychol-
ogy of pictorial representation (pp. 162–171). Ithaca, N.Y.: Cornell University
Press.
Gombrich, E. H. (1982). The image and the eye: Further studies in the psychology
of pictorial representation. Oxford: Phaidon.
Gombrich, E. H. (1983). Art and illusion: A study in the psychology of pictorial
representation. Oxford: Phaidon Press.
Gombrich, E. H. (1987). The art of the Greeks. In R. Woodfield (ed.), Reflections
on the history of art (pp. 11–17). Berkeley: University of California Press.
Goodman, N. (1968). The Languages of art. New York: Bobbs-Merrill.
Goodman, N. (1968/1976). Languages of art: An approach to a theory of symbols.
Indianapolis: Hackett.
Goodman, N. (1969). Languages of Art, 2d ed. Oxford: Oxford University Press.
Gorea, A., and Julesz, B. (1990). Context superiority in a detection task with line-
element stimuli: A low-level effect. Perception, 19, 5–16.
Gregory, R. L. (1977). Eye and brain: The psychology of seeing. London: Weiden-
feld and Nicolson.
387 References

Grüsser, O. J., and Landis, T. (1991). Visual agnosias and related disorders. In J.
Cronly-Dillon (ed.), Vision and visual dysfunction. Basingstoke, United Kingdom:
MacMillan.
Haber, R. N. (1980). Perceiving space from pictures: A theoretical analysis. In
M. A. Hagen (ed.), The perception of pictures, vol. 1 (pp. 3–31). New York:
Academic Press.
Hagen, M. A. (1985). There is no development in art. In N. H. Freeman and
M. V. Cox (eds.), Visual order: The nature and development of pictorial representa-
tion (pp. 59–77). Cambridge: Cambridge University Press.
Hagen, M. A. (1986). Varieties of realism: Geometries of representational art.
Cambridge: Cambridge University Press.
Hagen, M. A., and Giorgi, R. (1993). Where’s the camera? Ecological Psychology,
5, 65–84.
Hagen, M. A., and Jones, R. K. (1978). Cultural effects on pictorial perception:
How many words is one picture really worth? In R. D. Walk and H. L. Pick Jr.
(eds.), Perception and experience (pp. 171–212). New York: Plenum Press.
Harris, P. L., and Kavanaugh, R. D. (1993). Young children’s understanding of
pretense. Monographs of the Society for Research in Child Development, 58.
Harway, N. I. (1963). Judgment of distance in children and adults. Journal of
Experimental Psychology, 65, 385–390.
Hauck, G. (1879). Die subjektive Perspektive und die horizontalen Curvaturen des
dorischen Styls: Eine perspektivisch-ästhetische Studie (The subjective perspective
and the horizontal curvatures of the doric style: A perspectival aesthetic study).
Stuttgart, Germany: Wittwer.
Hawking, S. W. (1993). A brief history of time: From the Big Bang to black holes.
London: Bantam Press.
Hecht, H., van Doorn, A., and Koenderink, J. J. (1999). Compression of visual
space in natural scenes and in their photographic counterparts. Perception and
Psychophysics, 61, 1269–1286.
Heider, F. (1926). Ding und Medium. Symposium, 1, 109–157.
Helmholtz, H. von (1867). Handbuch der Physiologischen Optik. Hamburg: Voss.
Helmholtz, H. von (1894). Über den Ursprung der richtigen Deutung unserer
Sinneseindrücke (The origin and correct interpretation of our sense impressions).
Zeitschrift für Psychologie und Physiologie der Sinnesorgane, 7, 81–91.
Helmholtz, H. von (1896). Handbuch der physiologischen Optik, 2d ed. Hamburg:
Voss.
Helmholtz, H. von (1910). Handbuch der Physiologischen Optik. 3d ed., 3d vol.
Hamburg/Leipzig: Voss. (Expanded and coedited with A. Gullstrand, J. v. Kries,
and W. Nagel.)
Helmholtz, H. von (1911/1925). Treatise on physiological optics, 3d ed. (ed. and
trans. J. P. C. Southall). Rochester, N.Y.: Optical Society of America.
388 References

Hel-Or, Y., and Edelman, S. (1994). A new approach to qualitative stereo. In


S. Ullman and S. Peleg (eds.), Proceedings of the 12th international conference on
pattern recognition (pp. 316–320). Washington, D.C.: IEEE Press.
Hildebrand, A. (1893/1945). The Problem of Form in Painting and Sculpture
(trans. M. Meyer and R. M. Ogden). New York: G. E. Stechert. (Originally pub-
lished 1893, Straßburg.)
Hildebrand, A. (1908). Das Problem der Form in der bildenden Kunst. Strasbourg,
Germany: Heitz.
Hochberg. J. (1962). The psychophysics of picture perception. Audio-Visual Com-
munication Review, 10, 22–54.
Hochberg, J. (1978). Art and perception. In E. C. Carterette and M. P. Friedman
(eds.), Handbook of perception, vol. 10: Perceptual ecology (pp. 225–258). New
York: Academic Press.
Hochberg, J., and Brooks, V. (1962). Pictorial recognition as an unlearned ability:
A study of one child’s performance. American Journal of Psychology, 75, 624–628.
Hoffman, D. (2003). Colour and contour from apparent motion. In R. Maus-feld
and D. Heyer (eds.), Colour perception: From light to object. Oxford: Oxford
University Press.
Holway, A. H., and Boring, E. G. (1941). Determinants of apparent visual size
with distance variant. American Journal of Psychology, 54, 21–37.
Hopkins, R. (1998). Picture, image and experience: A philosophical inquiry. Cam-
bridge: Cambridge University Press.
Hopkins, R. (2000). Touching pictures, British Journal of Aesthetics, 40,
149–167.
Hopkins, R. (2002). Thomas Reid on Molyneux’s question. Manuscript in
preparation.
Horn, B. K. P., and Brooks, M. J. (1989). Shape from shading. Cambridge, Mass.:
MIT Press.
Hornbostel, E. M. von (1922). Über optische Inversion (About Optical Inversion).
Psychologische Forschung, 1, 130–156.
Huffman, D. A. (1971). Impossible objects as nonsense sentences. In B. Meltzer
and D. Mitchie (eds.), Machine Intelligence, vol. 6 (pp. 295–323). Edinburgh:
Edinburgh University Press.
Huizinga, J. (1938/1986). Homo Ludens. Boston: Beacon Press.
Hurlbert, A. (1998). Illusions and reality-checking on the small screen. Perception,
27, 633–636.
Husserl, E. (1980). Phantasie, Bildbewusstsein, Erinnerung. Zur Phänomenologie
der anschaulichen Vergegenwärtigungen. Text aus dem Nachlass (1898–1925)
(Phantasy, image consciousness, memory. On the phenomenology of intuitive
realizations. Text from the bequest estate 1898–1925), Husserliana, vol. 23, ed.
E. Marbach. Dordrecht, Netherlands: Kluwer.
389 References

Hyman, J. (1992). Perspective. In D. Cooper (ed.), A companion to aesthetics


(pp. 323–327). Oxford: Blackwell.
Ikeda, M., and Takeuchi, T. (1975). Influence of foveal load on the functional
visual field. Perception and Psychophysics, 18, 255–260.
Indow, T. (1991). A critical review of Luneburg’s model with regard to global
structure of visual space. Psychological Review, 98, 430–453.
Ittelson, W. H. (1952). The Ames demonstrations in perception. Princeton, N.J.:
Princeton University Press.
Ittleson, W. H. (1960). Visual space perception. New York: Springer.
Ittelson W. H. (1996). Visual perception of markings. Psychonomic Bulletin and
Review, 3, 171–187.
Jackendoff, R. (1987). Consciousness and the computational mind. Cambridge,
Mass.: MIT Press.
Jantzen, H. (1911). Die Raumdarstellung bei kleiner Augendistanz (Representa-
tion of space at near eye distance). Zeitschrift für Ästhetik und allgemeine Kunst-
wissenschaft, 6, 119–123.
Julesz, B. (1971). Foundations of cyclopean perception. Chicago: University of
Chicago Press.
Kaila, E. (1928). Gegenstandsfarbe und Beleuchtung (Object color and illumina-
tion). Psychologische Forschung, 3, 18–59.
Kalkofen, H. (1993). Szenengliederung und Szenenfolge im sogenannten Simultan-
bild (Scene segmentation and scene order in the so-called simulations image).
Paper presented at the 7th. Internationaler Kongreß der Deutschen Gesellschaft
für Semiotik, Tübingen, Germany.
Katz, D. (1911). Die Erscheinungsweisen der Farben und ihre Beeinflussung durch
die Individuelle Erfahrung. Leipzig, Germany: J. A. Barth.
Katz, D. (1935). The world of color. London: Kegan. (Translated in abridged form
from the second German edition, 1930.)
Kemp, M. (1990). The science of art: Optical themes in western art from
Brunelleschi to Seurat. New Haven, Conn.: Yale University Press.
Kennedy, J. M. (1974). A psychology of picture perception. San Francisco: Jossey-
Bass.
Kennedy, J. M. (1993). Drawing and the blind. New Haven, Conn.: Yale Univer-
sity Press.
Kennedy, J. M. (1997). How the blind draw. Scientific American, 276 (January),
pp. 60–65.
Kennedy, J. M. (in press). Drawings by Gaia, a blind girl. Perception.
Kennedy, J. M., and Bai, J. (2000). Cavanagh and Leclerc shape-from-shadow
pictures: Do line versions fail because of the polarity of the regions or the con-
tours? Perception, 29, 399–407.
390 References

Kennedy, J. M., and Domander, R. (1985). Shape and contour: The points of
maximum change are least useful for recognition Perception, 14, 367–370.
Kennedy, J. M., and Juricevic, I. (1999). Angular subtense and binocular vision:
Worse than worst. In M. A. Grealey and J. A. Thomson (eds.), Studies in percep-
tion and action, vol. 5. (pp. 24–27). Mahwah, N.J.: L. Erlbaum.
Kennedy, J. M., and Merkas, C. (2000). Depictions of motion devised by a blind
person. Psychonomic Bulletin and Review, 7, 700–706.
Kennedy, J. M., and Miyoshi, H. (1995). Combining illusions: A perfect additiv-
ity model fits imperfect additivity results. Paper presented at the meeting of the
Psychonomic Society, November 10–12, Los Angeles.
Kennedy, J. M., Nicholls, A., and Desrochers, M. (1995). From line to outline. In
C. Lange-Kuettner and G. V. Thomas (eds.), Drawing and looking (pp. 62–74).
Hemel Hempstead, United Kingdom: Simon and Schuster.
Kepes, G. (1944). The language of vision. Chicago: Paul Theobold.
Kersten, D., Bülthoff, H. H., Schwartz, B. L., and Kurtz, K. J. (1992). Interaction
between transparency and structure from motion. Neural Computation, 4, 573–589.
Kilpatrick, F. P., and Ittleson, W. H. (1953). The size-distance invariance hypoth-
esis. Psychological Review, 60, 223–231.
Klee, P. (1925/1981). Pädagogisches Skizzenbuch (Pedagogical sketch book). Mainz
and Berlin.
Klee, P. (1961). Notebooks (The thinking eye), vol. 1, ed. J. Spiller. London: Lund
Humphries.
Klein, F. (1893). Vergleichende Betrachtungen über neuere geometrische
Forschungen (the Erlanger Programm) (Comparative reflection on newer geomet-
rical research). Mathematische Annalen, 43, 63–100.
Knill, D. C., and Kersten, D. (1991). Apparent surface curvature affects lightness
perception. Nature, 351, 228–230.
Koenderink, J. J. (1990). Solid shape. Cambridge, Mass.: MIT Press.
Koenderink, J. J. (1998). Pictorial relief. Philosophical Transactions of the Royal
Society London, A, 356, 1071–1086.
Koenderink, J. J. (1999). Virtual psychophysics. Perception, 28, 669–674.
Koenderink, J. J., and van Doorn, A. J. (1995). Relief: Pictorial and otherwise.
Image and Vision Computing, 13, 321–331.
Koenderink, J. J., and Doorn, A. J. van (1997). The generic bilinear calibration-
estimation problem. International Journal of Computer Vision, 23, 217–234.
Koenderink, J. J., and van Doorn, A. J. (1998). The Structure of Relief. Advances
in Imaging and Electron Physics, 103, 65–150.
Koenderink, J. J., van Doorn, A. J., and Kappers, A. M. L. (1992). Surface per-
ception in pictures. Perception and Psychophysics, 52, 487–496.
Koenderink, J. J., van Doorn, A. J., and Kappers, A. M. L. (1994). On so-called
paradoxical monocular stereoscopy. Perception, 23, 583–594.
391 References

Koenderink, J. J., van Doorn, A. J., and Kappers, A. M. L. (1995). Depth relief.
Perception, 24, 115–126.
Koenderink, J. J., van Doorn, A. J., and Kappers, A. M. L. (1996). Pictorial
surface attitude and local depth comparisons. Perception and Psychophysics, 58,
163–173.
Koenderink, J. J., van Doorn, A. J., Kappers, A. M. L., and Todd, J. T. (2000a).
Ambiguity and the “mental eye” in pictorial relief. Investigative Ophthalmology
and Visual Science, 41, S36.
Koenderink, J. J., van Doorn, A. J., Kappers, A. M. L., and Todd, J. T. (2000b).
Directing the mental eye in pictorial perception. In B. E. Rogowitz and T. N. Pap-
pas (eds.), Proceedings of SPIE: Human vision and electronic imaging, 3959, 2–13.
Koenderink, J. J., van Doorn, A. J., Kappers, A. M. L., and Todd, J. T. (2001).
Ambiguity and the “mental eye” in pictorial relief. Percpetion, 30, 431–448.
Koenderink, J. J., Kappers, A. M. L., Pollick, F. E., and Kawato, M. (1997).
Correspondence in pictorial space. Perception and Psychophysics, 59, 813–827.
Koffka, K. (1935). Principles of Gestalt psychology. New York: Harcourt.
Koffka, K. (1953). Principles of Gestalt psychology. New York: Harcourt, Brace,
and World.
Kosslyn, S. (1994). Image and the Brain. Cambridge, Mass.: MIT Press.
Kraft, R., and Green, J. S. (1989). Distance perception as a function of photo-
graphic area of view. Perception and Psychophysics, 45, 459–466.
Kratzsch, K. (ed.)(1983). Illuminierte Holzschnitte der Luther-Bibel von 1534 (Illu-
minated wood cuts of the Luther-Bibel of 1534). Hanau, Germany: Dausien.
Kruskal, J. B. (1964). Multidimensional scaling by optimizing goodness of fit to
a nonmetric hypothesis. Psychometrika, 29, 707–719.
Kubovy, M. (1986). The Psychology of perspective and renaissance art. Cambridge:
University Press.
Kubovy, M., and Epstein, W. (in press). Internalization: A metaphor we can live
without. Behavioral and Brain Sciences, 24, 618–625.
Künnapas, T. (1968). Distance perception as a function of available visual cues.
Journal of Experimental Psychology, 77, 523–529.
Lakoff, G., and Johnson, M. (1999). Philosophy in the flesh. New York: Basic
Books.
Landy, M. S., Maloney, L. T., Johnston, E. B., and Young, M. (1995). Measure-
ment and modeling of depth cue combination: In defense of weak fusion. Vision
Research, 35, 389– 412.
Landy, M. S., Maloney, L. T., and Young, M. (1991). Psychophysical estimation
of the human depth combination rule. In P. S. Shenker (ed.), Proceedings of the
SPIE, 1383, Sensor fusion III: 3-D perception and recognition, 247–254.
Lee, M., and Bremner, G. (1987). The representation of depth in children’s draw-
ings of a table. Quarterly Journal of Experimental Psychology, 39A, 479–496.
392 References

Leeman, F., Elffers, J., and Schuyt, M. (1976). Hidden images. New York: Harry
N. Abrams.
Lenk, E. (1926). Über die optische Auffassung geometrisch-regelmässiger
Gestalten (On the optic prehension of geometric regular Gestalts). In F. Krueger
(ed.), Neue Psychologische Studien, vol. 1 (pp. 577–612). Munich, Germany:
Beck’sche Verlagsbuchhandlung.
Leonardo da Vinci (1989). Leonardo on Painting (trans. and ed. M. Kemp,
M. Walker, and M. Kemp). New Haven, Conn.: Yale University Press.
Leslie, A. M. (1987). Pretense and representation: The origins of “theory of mind.”
Psychological Review, 94, 412–426.
Lessing, G. E. (1964). Laokoon oder über die Grenzen der Malerei und Poesie
[Laokoon or on the limits of painting and poetry]. Originally published 1766.
Stuttgart, Germany: Recclam.
Levelt, W. J. M. (1984). Some perceptual limitation in talking about space. In
A. J. van Doorn, W. A. van de Grind, and J. J. Koenderink (eds.), Limits in per-
ception (pp. 323–358). Utrecht, Netherlands: VNU Science Press.
Levin, C. A., and Haber, R. N. (1993). Visual angle as a determinant of perceived
interobject distance. Perception and Psychophysics, 54, 250–259.
Lewis, D. (1969). Convention: A philosophical study. Oxford: Blackwell.
Lindberg, D. C. (1976). Theories of vision from Al-Kindi to Kepler. Chicago: Uni-
versity of Chicago Press.
Loomis, J. M., Da Silva, J. A., Philbeck, J. W., and Fukushima, S. S. (1996).
Visual perception of location and distance. Current Directions in Psychological
Science, 5, 72–77.
Loomis, J. M., and Philbeck, J. W. (1999). Is the anisotropy of perceived 3-D
shape invariant across scale? Perception and Psychophysics, 61, 397–402.
Lopes, D. M. M. (1992). Pictures, styles and purposes, British Journal of Aesthet-
ics, 32 (4), 330–341.
Lopes, D. M. M. (1996). Understanding pictures. Oxford: Clarendon Press.
Lopes, D. M. M. (1997). Art media and the sense modalities: Tactile pictures.
Philosophical Quarterly, 47, 425– 440.
Lübke, W., and Pernice, E. (1921). Die Kunst des Altertums (The art of antiquity).
Esslingen, Germany: Neff.
Luce, R. D., and Krumhansl, C. L. (1988). Measurement, scaling, and psy-
chophysics. In R. C. Atkinson, R. J. Herrnstein, G. Lindzey, and R. D. Luce
(eds.), Stevens’ handbook of experimental psychology (pp. 3–74). New York: Wiley.
Lumsden, E. A. (1980). Problems of magnification and minification: An explana-
tion of the distortions of distance, slant, and velocity. In M. Hagen (ed.), The per-
ception of pictures, vol. 1 (pp. 91–135). New York: Academic Press.
Luneburg, R. K. (1947). Mathematical analysis of binocular vision. Princeton, N.J.:
Princeton University Press.
393 References

Luo, R. C., and Kay, M. G. (1992). Data fusion and sensor integration: State-of-
the-art 1990s. In M. A. Abidi and R. C. Gonzalez (eds.), Data fusion in robotics
and machine intelligence (pp. 7–135). San Diego, Calif.: Academic Press.
Lyons, J. (1977). Semantics, 2 vols. Cambridge: University Press.
Lythgoe, J. N. (1979). The ecology of vision. Oxford: Clarendon Press.
Mach, E. (1903). Analyse der Empfindungen (The analysis of sensations). Jena,
Germany: Fischer.
Mackavey, W. R. (1980). Exceptional cases of pictorial perspective. In M. A.
Hagen (ed.), The perception of pictures, vol. 1 (pp. 213–223). New York: Academic
Press.
Margerie, A. de (ed.) (1994). Gustave Caillebotte, 1948–1894. Paris: Réunion des
Musées Nationaux.
Marler, P. (1999). On innateness: Are sparrow songs “learned” or “innate”?
In M. D. Hauser and M. Konishi (eds.), The design of animal communication
(pp. 293–318). Cambridge, Mass.: MIT Press.
Marr, D. (1982). Vision: A computational investigation into the human representa-
tion and processing of visual information. San Francisco: W.H. Freeman.
Marr, D., and Poggio, T. (1976). Cooperative computation of human stereo
vision. Science, 194, 283–287.
Massaro, D. W., and Cohen, M. M. (1993). The paradigm and the fuzzy logical
model of perception are alive and well. Journal of Experimental Psychology: Gen-
eral, 122, 115–124.
Maunsell, J., Sclar, G., Nealey, T., and DePriest, D. (1991). Extraretinal repre-
sentations in area V4 in the macaque monkey. Vision Neuroscience, 7, 651–573.
Mauries, P. (1997). Le trompe l’oeil. Paris: Gallimard.
Mausfeld, R. (2002). The physicalistic trap in perception theory. In D. Heyer and
R. Mausfeld (eds.), Perception and the physical world (pp. 75–112). Chichester,
United Kingdom: Wiley.
Mausfeld, R. (2003). The dual coding of colour: “Surface colour” and “illumina-
tion colour” as constituents of the representational format of perceptual primi-
tives. In R. Mausfeld and D. Heyer (eds.), Colour perception: From light to object.
Oxford: Oxford University Press.
Mausfeld, R., and Andres, J. (2002). Second-order statistics of colour codes mod-
ulate transformations that effectuate varying degrees of scene invariance and illu-
mination invariance. Perception, 31, 209–224.
Maynard, P. (1991). Real imaginings, Philosophy and Phenomenological Research,
51 (2), 389–394.
Maynard, P. (1994). The contest of meaning (book review). Journal of Aesthetics
and Art Criticism, 49 (4), 390–392.
Maynard, P. (1996). Perspective’s places. Journal of Aesthetics and Art Criticism,
54 (1), 23– 40.
394 References

Maynard, P. (1997a). The engine of visualization: Thinking through photography.


Ithaca, N.Y.: Cornell University Press.
Maynard, P. (1997b). Rev. of Damisch 1994. Journal of Aesthetics and Art Criti-
cism, 55 (1), 84–85.
Maynard, P. (2001). “The time it takes.” In J. Baetens (ed.), The Graphic Novel
(pp. 191–210). Louvain, Belgium: Louvain University Press.
Meltzoff, A. N. (1995). Understanding the intentions of others: Re-enactment of
intended acts by 18-month-old children. Developmental Psychology, 31, 838–850.
Meng, J. C., and Sedgwick, H. A. (2001). Distance perception mediated through
nested contact relations among surfaces. Perception and Psychophysics, 63, 1–15.
Meng, J. C., and Sedgwick, H. A. (2002). Distance perception across spatial dis-
continuities. Perception and Psychophysics, 64, 1–14.
Messaris, P. (1994). Visual literacy: Image, mind, and reality. Boulder, Colo.:
Westview Press.
Metz, C. (1974). Film language. New York: Oxford University Press.
Metzger, W. (1975). Gesetze des Sehens. Frankfurt/Main: Kramer.
Michalski, S. (1993). The Reformation and the visual arts. London: Routledge.
Michotte, A. (1948/1991). L’énigma psychologique de la perspective dans le dessin
linéaire. Bulletin de la Classe des Lettres de l’Académie Royale de Belgique, 34,
268–288. Republished as The psychological enigma of perspective in outline pic-
tures. In G. Thinès, A. Costall, and G. Butterworth (eds.), Michotte’s experimen-
tal phenomenology of perception (pp. 174–187). Hillsdale, N.J.: Erlbaum.
Michotte, A. (1954/1991). Autobiographie. Psychologica Belgica, 1, 190–217.
Republished as Autobiography. In G. Thinès, A. Costall, and G. Butterworth
(eds.), Michotte’s experimental phenomenology of perception (pp. 24–49). Hillsdale,
N.J.: Erlbaum.
Michotte, A. (1960/1991). Le réel et l’irréel dans l’image. Bulletin de la Classe des
Lettres de l’Académie Royale de Belgique, 46, 330–344. Republished as The real
and the unreal in the image. In G. Thinès, A. Costall, and G. Butterworth (eds.),
Michotte’s experimental phenomenology of perception (pp. 187–197). Hillsdale,
N.J.: Erlbaum.
Michotte, A., Thinès, G., and Crabbé, G. (1964/1991). Amodal completion of per-
ceptual structures. In G. Thinès, A. Costall, and G. Butterworth (eds.), Michotte’s
experimental phenomenology of perception (pp. 140–167). Hillsdale, N.J.: Erlbaum.
(French original appeared in 1964.)
Milbrath, C. (1998). Patterns of artistic development in children: Comparative
studies of talent. Cambridge: Cambridge University Press, 1998.
Mingolla, E., and Todd, J. T. (1986). Perception of solid shape from shading. Bio-
logical Cybernetics, 53, 137–151.
Miskie, D., Dainoff, M., Sherman, R., and Johnson, L. (1975). Does distance per-
ception change as the degree of enclosure changes? Some psychophysical studies
under real and simulated conditions. Man-Environment Systems, 5, 317–320.
395 References

Mitchell, W. J. (1992). The reconfigured eye: Visual truth in the post-


photographic era. Cambridge, Mass.: MIT Press.
Mitchison, G. J., and Westheimer, G. (1984). The perception of depth in simple
figures. Vision Research, 24, 1063–1073.
Moskowitz, B. A. (1978). The acquisition of language. Scientific American, 239,
82–96.
Nagata, S. (1991). How to reinforce perception of depth in single two-dimensional
pictures. In S. R. Ellis, M. K. Kaiser, and A. C. Grunwald (eds.), Pictorial com-
munication in virtual and real environments (pp. 527–545). London: Taylor and
Francis.
Nakayama, K., He, Z. J., and Shimojo, S. (1995). Visual surface representation:
A critical link between lower-level and higher-level vision. In S. M. Kosslyn and
D. N. Osherson (eds.), Visual cognition. An invitation to cognitive sciences, vol 2.
(pp. 1–70). Cambridge, Mass.: MIT Press.
Nakayama, K., Shimojo, S., and Ramachandran, V. S. (1990). Transparency:
Relation to depth, subjective contours, luminance, and neon color spreading.
Perception, 19, 497–513.
Nicholls, A. L., and Kennedy, J. M. (1992). Drawing development: From similar-
ity to direction. Child Development, 63, 227–241.
Nicholls, A. L., and Kennedy, J. M. (1995). Foreshortening in cube drawings by
children and adults. Perception, 24, 1443–1456.
Nijhawan, R. (1997). Visual decomposition of colour through motion extrapola-
tion. Nature, 386, 66–69.
Palmer, S. E. (1999). Vision Science: Photons to Phenomenology. Cambridge,
Mass.: MIT Press.
Panofsky, E. (1924/1925). Die Perspektive als “symbolische Form.” In Vorträge
der Bibliothek Warburg. Leipzig and Berlin. (Translated in New York University
samizdat typescript, n.d.)
Panofsky, E. (1927/1991). Die Perspektive als “symbolische Form.” In Vor-
träge der Bibliothek Warburg 1924/25 (pp. 258–330). Leipzig and Berlin. Re-
published as Perspective as symbolic form (trans. C. S. Wood). New York: Zone
Books.
Peacocke, C. (1987). Depiction. Philosophical Review, 96, 383–410.
Peacocke, C. (1989). Perceptual content. In J. Lamog, J. Perry, and H. Wettstein
(eds.), Themes from Kaplan. New York: Oxford University Press.
Pearson, D., Hanna, E., and Martinez, K. (1990). Computer generated cartoons.
In H. Barlow, C. Blakemore, and M. Weston-smith (eds.), Images and under-
standing (pp. 46–50). Cambridge: Cambridge University Press.
Petersen, M. A. (1999). What’s in a stage name? Comment on Vecera and O’Reilly
(1998). Journal of Experimental Psychology: Human Perception and Performance,
25, 276–286.
Pinker, S. (1997). How the mind works. New York: W. W. Norton.
396 References

Pirenne, M. H. (1952/1953). The scientific basis of Leonardo da Vinci’s theory of


perspective. British Journal for the Philosophy of Science, 3, 169–185.
Pirenne, M. H. (1970). Optics, painting and photography. Cambridge: Cambridge
University Press.
Pirenne, M. H. (1975). Vision and art. In E. C. Carterette and M. P. Friedman
(eds.), Handbook of perception, vol. 5: Seeing (pp. 434–490). New York: Academic
Press.
Polanyi, M. (1970a). What Is a painting? The British Journal of Aesthetics, 10,
225–236.
Polanyi, M. (1970b). What Is a painting? American Scholar, 49, 655–669.
Pollick, F. E., Giblin, P. J., Rycroft, J., and Wilson, L. L. (1992). Human recov-
ery of shape from profiles. Behaviormetrika, 19, 65–79.
Posner, M. I., and Raichle, M. E. (1994). Images of mind. Scientific American
Library. New York: Freeman & Co.
Prazdny, K. (1986). Three-dimensional structure from long-range apparent
motion. Perception, 15, 619–625.
Racinet, A. (1888/1995). Weltgeschichte der Kostüme (World history of costumes).
Cologne: Parkland.
Ramachandran, V. S. (1986). Capture of stereopsis and apparent motion by illu-
sory contours. Peception and Psychophysics, 39, 361–373.
Ramachandran, V. S. (1990). Interactions between motion, depth, color, and form:
The utilitarian theory of perception. In C. Blakemore (ed.), Vision: Coding and
efficiency (pp. 346–360). Cambridge: Cambridge University Press.
Ramachandran, V. S., and Hirstein, W. (1999). The science of art: A neurological
theory of aesthetic experience. Journal of Consciousness Studies, 6, 15–51.
Ratoosh, P. (1949). On interposition as a cue for the perception of distance.
Proceedings of the National Academy of Sciences, Washington, 35, 257–259.
Rehkämper, K. (1995). Analoge Repräsentationen (Analog representations). In
K. Sachs-Hombach (ed.), Bilder im Geiste. Zur kognitiven und erkenntnistheoretis-
chen Funktion piktorialer Repräsentationen (Mental images: About the cognitive
and epistemological function of pictorial representation) (pp. 63–105). Amsterdam
and Atlanta, Ga.: Rodopi.
Reid, T. (1764). An inquiry into the human mind, on the principles of common sense.
Edinburgh: A. Kincaid and J. Bell.
Reisberg, D., and Chambers, D. (1991). Neither pictures nor propositions: What
can we learn from a mental image? Canadian Journal of Psychology, 45, 336–352.
Richter, J. P. (ed., and trans.) (1883). The literary works of Leonardo da Vinci.
London: Samson, Low, Marston, Searle, and Rivington. Reprinted as The note-
books of Leonardo da Vinci. New York: Dover Press, 1970.
Riedl, R. (1988). Biologie der Erkenntnis. Munich: Deutscher Taschenbuch
Verlag.
397 References

Rock, I. (1983). The Logic of perception. Cambridge, Mass.: MIT Press.


Rock, I. (1984). Perception. New York: Scientific American Books.
Rogers, B. J., and Collett, T. S. (1989). The appearance of surfaces specified by
motion parallax and binocular disparity. The Quarterly Journal of Experimental
Psychology, 41A, 697–717.
Rogers, S. (1995). Perceiving pictorial space. In W. Epstein and S. Rogers (eds.),
Handbook of perception and cognition, 2d ed., vol. 5: Perception of space and
motion (pp. 119–163). San Diego, Calif.: Academic Press.
Rogers, S. (1996). The horizon-ratio relation as information for relative size in
pictures. Perception and Psychophysics, 58, 142–152.
Rogers, S. (1997). Convention and constraint in pictorial perspective. In G. L.
Torenvliet and K. J. Vicente (eds.), Perception and action. Proceedings of the ninth
international conference on perception and action, (p. 112). Mahwah, N.J.:
L. Erlbaum.
Rogers, S. (2000). The emerging concept of information. Ecological Psychology,
12 (4), 335–343.
Rogers, S. (in press). Through Alice’s glass: The creation and perception of other
worlds in movies, pictures and virtual reality. In J. D. Anderson and B. F. Ander-
son (eds.), Moving image theory: Ecological considerations. Carbondale IL: South-
ern Illinois University Press.
Rogers, S., and Costall, A. (1983). On the horizon: Picture perception and Gibson’s
concept of information. Leonardo, 16 (3), 180–182.
Rogers, S., and Fichandler, C. E. (2000). Eye-height scaling of layout in pictures,
trompe l’oeil, and real scenes. Abstracts of the Psychonomic Society, 5, 39.
Rogers, S., and Pollard, W. (1998). Perceived size and distance in real scenes and
in photographs. Abstracts of the Psychonomic Society, 3, 59.
Rogers, S., and Watson, J. G. (1996a). Horizon-based relative size judgments in
pictures, movies and virtual reality. Investigative Ophthalmology and Visual Sci-
ence, 37 (3), S519.
Rogers, S., and Watson, J. G. (1996b). The horizon-ratio: Size-distance invariance
revisited. Abstracts of the Psychonomic Society, 1, 25.
Rubin, E. (1915). Synsoplevede figurer (Visually perceived figures). Copenhagen:
Gyldendals.
Sachs, H. (1990). Isotrope Geometrie des Raumes (Isotropic geometry of space).
Brunswick, Germany: Friedr. Vieweg and Sohn.
Sachs-Hombach, K. (1999). Bilder als wahrnehmungsnahe Zeichen. In J. Mittel-
strass (ed.), Die Zukunft des Wissens (The future of knowledge) (pp. 1351–1358).
Constance, Germany: Universitätsverlag Konstanz.
Sachs-Hombach, K., and Rehkämper, K. (1998). Thesen zu einer Theorie
bildhafter Darstellung. In K. Sachs-Hombach and K. Rehkämper (eds.),
Bild—Bildwahrnehmung—Bildverarbeitung. Interdisziplinäre Beiträge zur Bild-
wissenschaft (pp. 119–124). Wiesbaden, Germany: Deutscher Universitätsverlag.
398 References

Salt, B. (1983). Film style and technology: History and analysis. London: Starword.
Sanders, A. F. (1970). Some aspects of the selective process in the functional
visual field. Ergonomics, 13, 101–117. (Cited in Ikeda and Takeuchi 1975).
Sartre, J.-P. (1940). L’imaginaire. Psychologie phénoménologique d’imagination
(The imaginary: Phenomenal psychology of imagination). Paris: Gallimard.
Scharf, A. (1968). Art and photography. London: Penguin.
Schier, F. (1986). Deeper into pictures. Cambridge: Cambridge University Press.
Scholz, O. R. (1991): Bild, Darstellung, Zeichen. Philosophische Theorien bildhafter
Darstellung (Image, representation, symbol: Philosophical theories of pictorial
representation). Freiburg and Munich: Alber.
Schöne, W. (1954). Über das Licht in der Malerei (On the light in painting). Berlin:
Gebr. Mann.
Schriever, W. (1925). Experimentelle Studien über stereoskopisches Sehen (Exper-
imental studies of stereoscopic vision). Zeitschrift für Psychologie, 96, 113–170.
Schuster, D. H. (1964). A new ambiguous figure. American Journal of Psychology,
77, 673.
Schwartz, B. J., and Sperling, G. (1983). Luminance controls the perceived 3-D
structure of dynamic 2-D displays. Bulletin of the Psychonomic Society, 21,
456–458.
Schwartz, R. (1994). Vision: Variations on some Berkelian themes. Cambridge, MA:
Blackwell.
Schwartz, R. (1996). Directed Perception. Philosophical Psychology, 9, 81–91.
Searle, J. R. (1993). Metaphor. In A. Ortony (ed.), Metaphor and thought, 2d ed.
(pp. 83–111). Cambridge: Cambridge University Press.
Sedgwick, H. A. (1973). The visible horizon: A potential source of visual infor-
mation for the perception of size and distance. Doctoral dissertation, Cornell
University. Dissertation Abstracts International, 34, 1301B–1302B. (University
Microfilms No. 73–22,530.)
Sedgwick, H. A. (1980). The geometry of spatial layout in pictorial representa-
tions. In M. A. Hagen (ed.), The Perception of Pictures, vol. 1 (pp. 33–90). New
York: Academic Press.
Sedgwick, H. A. (1983). Environment-centered representation of spatial layout:
Available visual information from texture and perspective. In J. Beck, B. Hope,
and A. Rosenfeld (eds.), Human and machine vision (pp. 425–458). New York:
Academic Press.
Sedgwick, H. A. (1986). Space perception. In K. R. Boff, L. Kaufman, and J. P.
Thomas (eds.), Handbook of perception and performance, vol. 1: Sensory process-
es and perception pp. 21.1–21.57. New York: Wiley.
Sedgwick, H. A. (1987a). Layout2: A production system modeling visual perspec-
tive information. In Proceedings of the IEEE First International Conference on
Computer Vision (pp. 662–666). Washington, DC: IEEE Computer Society Press.
399 References

Sedgwick, H. A. (1987b). A production system modeling high-level visual per-


spective information for spatial layout. Technical Report no. 298, New York Uni-
versity Department of Computer Science.
Sedgwick, H. A. (1989). Combining multiple forms of visual information to spec-
ify contact relations in spatial layout. In P. S. Schenker (ed.), Proceedings of SPIE,
vol. 1198: Sensor fusion II: Human and machine strategies (pp. 447–458).
Bellingham, Wash.: SPIE.
Sedgwick, H. A. (1991). The effects of viewpoint on the virtual space of pictures.
In S. R. Ellis (ed.), Pictorial communication in virtual and real environments,
pp. 460– 479. New York: Taylor and Francis.
Sedgwick, H. A. (2001). Visual space perception. In E. B. Goldstein (ed.), Black-
well handbook of perception (pp. 128–167). Oxford: Blackwell.
Sedgwick, H. A., and Gillam, B. (submitted). A non-modular approach to visual
space perception.
Sedgwick, H. A., and Levy, S. (1985). Environment-centered and viewer-centered
perception of surface orientation. Computer Vision, Graphics, and Image Process-
ing, 31, 248–260.
Sedgwick, H. A., and Nicholls, A. L. (1993a). Cross talk between the picture sur-
face and the pictured scene: Effects on perceived shape. Perception, 22 (supple-
ment), 109.
Sedgwick, H. A., and Nicholls, A. L. (1993b). Interaction between surface and depth
in the Ponzo illusion. Investigative Ophthalmology and Visual Science, 34, 1184.
Sedgwick, H. A., and Nicholls, A. L. (1994). Distortions of pictorial space pro-
duced by optical minification. Investigative Ophthalmology and Visual Science, 35,
2111.
Sedgwick, H. A., and Nicholls, A. L. (1996). Picture perception in children: Inter-
actions between the picture surface and the depicted scene. Investigative Ophthal-
mology and Visual Science, 37, S518.
Sedgwick, H. A., Nicholls, A. L., and Brehaut, J. (1995). Perceptual interaction of
surface and depth in optically minified picture. Investigative Ophthalmology and
Visual Science, 36, S667.
Sedgwick, H. A., Placide, F., Raul, C., and Flagg, T. (1991). Perceptual cross talk
between pictured spatial layout and the picture surface. Proceedings of the Seventh
Annual Meeting of the International Society for Psychophysics, Durham, N.C.,
October 17–21.
Shepard, R. N. (1980). Multidimensional scaling, tree-fitting, and clustering.
Science, 210, 390–398.
Shepard, R. N. (1990). Mindsights. New York: W. H. Freeman.
Smith, R. (1974). Natural versus scientific vision: the foreshortened figure in the
Renaissance. Gazette des Beaux-arts, 84, 239–248.
Speed, H. (1913). The practice and science of drawing. London: Seeley, Service, and
Company.
400 References

Spencer, J. R. (1966). Leon Battista Alberti on Painting. New Haven, Conn.: Yale
University Press.
Spivak, M. (1975). Differential geometry, vol. 3. Berkeley, CA: Publish or perish,
Inc.
Steinberg, S. (1966). Le masque. Paris: Maeght Editeur.
Stemberger, G. (1990). 2000 Jahre Christentum–Illustrierte Kirchengeschichte in
Farbe (2000 years of Christianity—Illustrated ecclesiastic history in color). Erlan-
gen, Germany: Karl Müller.
Stern, W. (1930). Psychology of early childhood. London: George Allen and Unwin.
Stevens, K. A., and Brookes, A. (1988). Integrating stereopsis with monocular
interpretations of planar surfaces. Vision Research, 28, 371–386.
Stevens, K. A., Lees, M., and Brookes, A. (1991). Combining binocular and
monocular curvature features. Perception, 20, 425–440.
Stevens, S. S. (1951). Mathematics, measurement, and psychophysics. In S. S.
Stevens (ed.), Handbook of Experimental Psychology (pp. 1–49). New York: Wiley.
Stillings, N., Feinstein, M., Garfield, J., Rissland, E., Rosenbaum, D., Weisler, S.,
and Baker-Ward, L. (1987). Cognitive Science: An Introduction. Cambridge, Mass.:
MIT Press.
Strieder, P. (ed.) (1996). Dürer. Augsburg, Germany: Bechtermünz.
Strubecker, K. (1956). Einführung in die höhere Mathematik mit besonderer
Berücksichtigung ihre Anwendungen auf Geometrie, Physik, Naturwissenschaften
und Technik (Introduction into higher mathematics with particular emphasis on
its application to geometry, physics, natural science and technology), vol. 1:
Grundlagen (Foundations). Munich: R. Oldenburg Verlag.
Swedlund, C. (1981). Photography: A handbook of history, materials, and pro-
cesses, 2d ed. New York: Holt, Rinehart, and Winston.
Talbot, W. H. F. (1844–1846/1969). The pencil of nature, unpaginated reprint.
New York: Da Capo.
Taylor, B. (1719). New principles of linear perspective: Or the art of designing on a
plane the representations of all sorts of objects, in a more general and simple method
than has been done before. London: Printed for R. Knaplock at the Bishop’s Head
in St. Paul’s Churchyard.
Teghtsoonian, R. (1973). Range effects in psychophysical scaling and a revision
of Stevens’ law. American Journal of Psychology, 86, 3–27.
Ten Doesschate, G. (1964). Perspective: Fundamentals, controversials, history.
Nieuwkoop, Netherlands: B. de Graaf.
Te Peerdt, E. (1899). Das Problem der Darstellung des Momentes der Zeit in
den Werken der malenden und zeichnenden Kunst (The problem of representing a
moment in time in paintings and drawings). Strasbourg, Germany: Heitz.
Terzopoulos, D. (1986). Integrating visual information from multiple sources. In
A. Pentland (ed.), From pixels to predicates (pp. 111–142). Norwood, N.J.: Ablex.
401 References

Thiéry, A. (1895). Ueber geometrisch-optische Täuschungen (On geometric-optical


illusions). Philosophische Studien, 11, 307–337.
Todorovic, D. (2002). Comparative overview of perception of distal and proximal
visual attributes. In D. Heyer and R. Mausfeld (eds.), Perception and the physical
world (pp. 37–74). New York: Wiley.
Tomasello, M., and Call, J. (1997). Primate cognition. Oxford: Oxford University
Press.
Topper, D. (2000). On anamorphosis: Setting some things straight. Leonardo, 33
(2), 115–124.
Trueswell, J. C., and Hayhoe, M. M. (1993). Surface segmentation mechanisms
and motion perception. Vision Research, 33, 313–328.
Turhan, M. (1937). Über räumliche Wirkungen von Helligkeitsgefällen (About
spatial impressions of contrast decrements). Psychologische Forschung, 21, 1–49.
Van Essen, D., and Anderson, C. A. (1990). Information processing strategies and
the pathways in the primate retina and visual cortex. In S. F. Zornetzer, J. L.
Davis, and C. Lau (eds.), Introduction to neural and electronic networks (pp.
43–52). Orlando, FL: Academic Press.
Van Essen, D., Anderson, C. A., and Olshausen, B. A. (1994). Dynamic routing
strategies in sensory, motor, and cognitive processing. In C. Koch and J. L. Davis
(eds.), Large scale neuronal theories of the brain (pp. 271–299), MIT Press.
Varnedoe, J. K. (1990) A fine disregard: What makes modern art modern. New
York: Abrams.
Victoria, J. (1982). Correspondence between implied points of view and selected
points of view in children’s drawings of familiar and unfamiliar objects. Review
of Research in Visual Arts Education, 15, 33–42.
Vitz, P. C. (1979). Visual science and modernist art: Historical parallels. In C. F.
Nodine and D. F. Fisher (eds.), Perception and pictorial presentation (pp. 134–166).
New York: Praeger.
Vollmer, G. (1990). Evolutionäre Erkenntnistheorie (Evolutionary epistemology).
Stuttgart, Germany: S. Hirzel.
Wade, N. J., and Hughes, P. (1999). Fooling the eyes: Trompe l’oeil and reverse
perspective. Perception, 28, 1115–1119.
Wagner, M. (1985). The metric of visual space. Perception and Psychophysics, 38,
483–495.
Wallach, H., and Karsh, E. B. (1963). The modification of stereoscopic depth-
perception and the kinetic depth-effect. American Journal of Psychology, 76,
429–435.
Wallach, H., Moore, M. E., and Davidson, L. (1963). Modification of stereo-
scopic depth-perception. American Journal of Psychology, 76, 191–204.
Wallach, H., Weisz, A., and Adams, P. A. (1956). Circles and derived figures in
rotation. American Journal of Psychology, 69, 48–59.
402 References

Walton, K. (1984). Transparent pictures: On the nature of photographic realism.


Critical Inquiry, 11, 246–277.
Walton, K. (1990). Mimesis as make-believe. Cambridge, Mass.: Harvard Univer-
sity Press.
Waltz, D. (1975). Understanding line drawings of scenes with shadows. In P. H.
Winston (ed.), The psychology of computer vision (pp. 19–91). New York:
McGraw-Hill.
Ware, W. R. (1900). Modern perspective, rev. ed. New York: Macmillan.
Wehner, R. (1987). “Matched filters”—neural models of the external world. Jour-
nal of Comparative Physiology A, 161, 511–531
Weisstein, N., and Harris, C. S. (1974). Visual detection of line segments: An
object superiority effect. Science, 186, 752–755.
Weisstein, N., and Wong, E. (1986). Figure-ground organization and the spatial
and temporal responses of the visual system. In E. C. Schwab and H. C. Nusbaum
(eds.), Pattern recognition by humans and machines, vol. 2: Visual perception
(pp. 31–64) Orlando, Fla.: Academic Press.
Wheeler, M. E., Petersen, S. E., and Buckner, R. L. (2000). Memory’s echo: vivid
remembering reactivates sensory-specific cortex. Proceedings of the National Acad-
emy of Sciences, 97, 11125–11129.
White, J. (1967). The birth and rebirth of pictorial space. London: Faber and Faber.
White, J. (1987). The birth and rebirth of pictorial space, 3d ed. London: Faber.
Wickhoff, F. (1912). Römische Kunst (Die Wiener Genesis) (Roman art [the
Viennese Genesis]). In M. Dvorak (ed.), Die Schriften Franz Wickhoffs (The writ-
ing of Franz Wickhoffs) (pp. 1–224). Berlin.
Wiesing, L. (2000). Phänomene im Bild (Phenomena in pictures). Munich: Fink.
Willats, J. (1977). How children learn to draw realistic pictures. Quarterly Journal
of Experimental Psychology, 29, 367–382.
Willats, J. (1992a). The representation of extendedness in children’s drawings of
sticks and discs. Child Development, 63, 692–710.
Willats, J. (1992b). Seeing lumps, sticks and slabs in silhouettes. Perception, 21,
481– 496.
Willats, J. (1997). Art and representation: New principles in the analysis of pictures.
Princeton, N.J.: Princeton University Press.
Wimmer, H., and Perner, J. (1983). Beliefs about beliefs: Representation and con-
straining functions of wrong beliefs of young children’s understanding of decep-
tion. Cognition, 13, 103–128.
Wittkower, R. (1977). Sculpture, processes and principles. New York: Harper and
Row.
Wollheim, R. (1977). Representation: The philosophical contribution to psychol-
ogy. In G. Butterworth (ed.), The Child’s representation of the world (pp. 173–188).
New York: Plenum Press.
403 References

Wollheim, R. (1980). Seeing-as, Seeing-in, and Pictorial Representation. In Art


and its Objects, 2d ed. (pp. 205–226). Cambridge: Cambridge University Press.
Wollheim, R. (1973). On art and the mind. London: Allen Lane.
Wollheim, R. (1987). Painting as an art. London: Thames and Hudson.
Wood, C. S. (1995). Review of Damisch 1994, Art Bulletin, 77 (December),
677–682.
Wundt, W. (1910). Grundzüge der Physiologischen Psychologie (The basics of
physiological psychology), 6th rev. ed., vol. 2. Leipzig, Germany: Engelmann.
Wyszecki, G., and Stiles, W. S. (1967). Color science: Concepts and methods, quan-
titative data and formulas. New York: John Wiley and Sons.
Yaglom, I. M. (1979). A simple non-euclidean geometry and its physical basis: An
elementary account of Galilean geometry and the Galilean principle of relativity
(trans. from Russian by A. Shenitzer). Springer, N.Y.: Heidelberg Science Library.
Yaglom, I. M. (1988). Felix Klein and Sophus Lie: Evolution of the idea of symme-
try in the nineteenth century. (Trans. from Russian by S. Sossinsky, eds. H. Grant
and A. Shenitzer). Boston: Birkhäuser.
Yang, T., and Kubovy, M. (1999). Weakening the robustness of perspective: Evi-
dence for a modified theory of compensation in picture perception. Perception and
Psychophysics, 61, 456– 467.
Yellott, J. I. (1981). Binocular depth inversion. Scientific American, 245, 148–159.
Youngs, W. M. (1976). The influence of perspective and disparity cues on the per-
ception of slant. Vision Research, 16, 79–82.
Zacks, J. M., and Tversky, B. (2001). Event structure in perception and concep-
tion. Psychological Bulletin, 127, 3–21.
Zeki, S. (1993). A Vision of the brain. Oxford: Blackwell.
This page intentionally left blank
Contributors

Margaret Atherton Dieter Heyer


Department of Philosophy Martin-Luther-Universität Halle-
University of Wisconsin at Wittenberg
Milwaukee Institut für Psychologie
Milwaukee, Wisconsin Halle, Germany
Juan Bai Robert Hopkins
University of Toronto Department of Philosophy
Department of Psychology University of Birmingham
Scarborough, Ontario, Canada Birmingham, United Kingdom
James E. Cutting Igor Juricevic
Department of Psychology Department of Psychology
Cornell University University of Toronto
Ithaca, New York Scarborough, Ontario, Canada
Andrea van Doorn Hermann Kalkofen
Helmholtz Instituut—Buys Ballot Universität Göttingen
Laboratory Georg-Elias-Müller Institut für
Universiteit Utrecht Psychologie
Utrecht, Netherlands Göttingen, Germany
Heiko Hecht John M. Kennedy
Department of Experimental Department of Psychology
Psychology University of Toronto
Johannes Gutenberg-Universität Scarborough, Ontario, Canada
Mainz
Mainz, Germany
406 Contributors

Jan J. Koenderink Klaus Sachs-Hombach


Helmholtz Instituut, Buys Ballot Otto-von-Guericke-Universität
Laboratory Magdeburg
Universiteit Utrecht Interdisziplinäre Forschungsstelle
Utrecht, Netherlands für Computervisualistik Institut
für Simulation und Graphik
Rainer Mausfeld
Magdeburg, Germany
Christian-Alberchts-Universität zu
Kiel Robert Schwartz
Institut für Psychologie Department of Philosophy
Kiel, Germany University of Wisconsin at
Milwaukee
Patrick Maynard
Milwaukee, Wisconsin
Department of Philosophy, Talbot
College H. A. Sedgwick
University of Western Ontario Department of Vision Sciences
London, Ontario, Canada SUNY College of Optometry
New York, New York
Reinhard Niederée
Christian-Alberchts-Universität zu John Willats
Kiel The Gatehouse
Institut für Psychologie Kingweston
Kiel, Germany Somerton, Somerset, United
Kingdom
Klaus Rehkämper
Carl von Ossietzky Universität Richard Wollheim
Oldenburg Department of Philosophy
Institut für Philosophie University of California at
Oldenburg, Germany Berkeley
Berkeley, California
Sheena Rogers
Department of Psychology
James Madison University
Harrisonburg, Virginia
Mark Rollins
Department of Philosophy
Washington University
Saint Louis, Missouri
Index

Action space, 224 tallying effect of, 8


Alain, Daniel, 376 T-junction and, 139–141
Alberti, L. B., 127, 149 tone and, 134–135
Alberti window, 20, 26, 39–40 trompe-l’oeil pictures and, 5, 82, 96, 172
Alhazen, 65, 67 visibility limits and, 9–13
Ambassadors (Holbein), 95 Art and Illusion (Gombrich), 4, 367
Ames room, 22, 52, 60, 77 Arti del disegno, 208
Anderson, Charles, 113, 115 Atherton, Margaret, 405
Another Camel! (Klee), 131 Atkins, Joseph, 35
Arcimboldo, Givseppe, 98
Arntz, Gerd, 44 Bai, Juan, xviii, 321–354, 405
Art, xii–xiii. See also Pictorial space Baj, Enrico, 50
Alberti on, 149 Baxandall, Michael, 208
Byzantine, 131–132 Bayesian model, 22
children and, 129–130, 136–137 Bazin, André, 319
conjoint representations and, 25–56 (see Beholder’s share, 277–278
also Conjoint representations) Berger, John, 195, 202
construction rules for, 20 Bergman, Richard, 221
dimension and, 148–149 Berkeley, George, 88, 327
drawing and, 133–134, 149–150 (see also Big Fishes Eat the Small Fishes, The
Drawing) (Brueghel the Elder), 44
Greek, 130 Birds, The (Hitchcock), 195–196
Hagen on, 128–131 Birth of Saint John the Baptist (de Paolo),
Huffman on, 137, 139 132
imprecision and, 151–156 Bischoff, Michael, 360
imputed mental states and, 46–51 Blickfeld, 356, 363
layout and, 231–236 Borders, 321
Marr on, 133 affinity and, 331–336
optic array and, 135 children and, 346
Oriental, 128 contour and, 348–353
perspective fallacies and, 203–207 discontinuities and, 346–348
postmodernism and, 194–195 grouping and, 348–349
pretense play and, 45–46 induction and, 324–326
relativism and, 204 information and, 331–336
Renaissance and, 19–20 kinds of visual, 328–336
representational content and, 3–13 number of, 336
resemblance view and, 150–165 stereo, 353–354
seeing-in and, 4–15 texture gradients and, 331–332
spectator’s education and, 193–195 Bosch, Hieronymus, 44
surfaces and, 321–354 (see also Surfaces) Botticelli, Sandro, 375
symbolic rules and, 128–143 Brancusi, Constantin, 275
408 Index

Braunstein, Myron, 72 Computational Gibsonism, 99


Breakfast (Gris), 137–138, 141 Computers, 71–72
Brightness, 38–40. See also Color; Light Congruence groups, 279–287
optic array and, 135 Conjoint representations, 18–19, 21, 57–60
proximal mode and, 54–56 cognitive architecture and, 36–51
vagueness and, 51–56 color and, 52–54
Brookes, Allen, 23–24 continuous transition and, 26
Bühler, Karl, 39, 366 imputed mental states and, 46–51
interlocked nature of, 28, 30–36
Cameras. See Photography language/meaning and, 41–45
Campo di Santi Giovanni e Paolo parameter setting and, 30–36
(Canaletto), 126–127, 134 pretense play and, 45–46
Canaletto, Giovanni, 126–127, 134 proximal mode and, 54–56
Cartesian projection, 207, 210n5, 257, 274, psychophysics and, 37–41
281 vagueness and, 51–56
Cartoon operator, 135 Construction (station) point, 202, 206, 211n7
Cayley-Klein geometries, 283–284 Constructivism, 99, 102
Cennini, Cennino, 127 Contour
Chagall, Marc, 132 ambiguity of, 349–353
Chatman, Seymour, 364–366 discontinuities and, 346–348
Chishty, Rizvana, 339 outline and, 348
Chomsky, Noam, 31, 50–51, 137 Control mechanisms, 116
Chromatic illumination, 39 Convention, 105–107
Churchland, Paul, 112–114, 116–117, 119 dimension and, 147–148
Cinerama movies, 62 experience and, 160–161
Clowes, Max, 137 imprecision and, 151–156
Cognition, 57–60 modularity thesis and, 108–110
cross-talk and, 71–74 nonconceptual spatial perception and,
figure-ground ambiguity and, 41 110–113
imputed mental states and, 46–51 optical laws vs. symbolic rules, 125–143
language/meaning and, 41–45 pictorial space and, 113–118
nonconceptual spatial perception and, resemblance view and, 150–165, 167–177
110–113 seeing-in and, 147
parameter setting and, 30–36 spectator’s education and, 157–158
perceived depth and, 20–21 theory of pictures and, 146–149
pictorial relief and, 259–260 validity and, 145
pretense play and, 45–46 variety accommodation and, 149–156
proximal mode and, 54–56 Cornelius, Hans, 357–360
psychophysics and, 37–41 Costall, Alan, 218
reinforcement and, 117 Cross-talk, 91
representational primitives and, 32–36 compensation hypothesis and, 94–96
vagueness and, 51–56 spatial layout and, 71–74
Color Cue integration, 21ff
chromatic illumination and, 39 Cueing, 93, 372
discontinuities and, 346–348 cognitive architecture and, 36–51
induction and, 326 color and, 38–40
optic array and, 135 conjoint representations and, 25–56 (see
planar aspects of, 81–82 also Conjoint representations)
proximal mode and, 54–56 cross-talk and, 71–74, 91, 94–96
representational primitives and, 36, 38–40 depth/spatial conflicts and, 21–25
resemblance view and, 150–165 dimension and, 37–38
simultaneous location and, 39 dual perception and, 89–92
tone and, 134–135 illusion and, 24
vagueness and, 51–56 integration, 89–92
Commonsense taxonomies, 17–19 objects and, 40–41
Compensation hypothesis, 94–96 parameter setting and, 30–36
409 Index

pictorial relief and, 251, 266–267, 280– children and, 324–325, 346
281 contour ambiguity and, 349–353
psychophysics and, 37–38 curved lines and, 336–337
reality and, 27–28 discontinuities and, 346–348
segregation, 89–92 element concentration and, 337–340
surfaces and, 40–41 foreshortening and, 340–345
triggering and, 30–36 induction and, 324–327
vagueness and, 51–56, 92 layout and, 231–236
Curvilinearity, 181–182 line junctions and, 337
Cutting, James E., xvii, 104, 117, 215–238, optical edge and, 349–353
405 outline and, 324–326
Dual perception
Danto, Arthur, 107, 109–110 attentional shift and, 81
da Vinci, Leonardo, 180, 225, 247 conjoint representations and, 25–56 (see
de Barde, Alexandre I. Leroy, 7 also Conjoint representations)
De Bry, Theodor, 369, 371, 373 content strategies and, 99–122
Della Pittura (Alberti), 127 convention and, 105–113
Denis, Maurice, 247 cross-talk and, 91
de Paolo, Giovanni, 132 cueing and, 89–92
Depth. See Dimension dissociation and, 96–97
Descartes, Rene, 65 ecological specificity and, 101
Dimension, 211n11, 319 equivalence and, 88–89
absolute, 308–311 functional aspects of, 92–94
art and, 148–149 imagination and, 78
borders and, 353–354 (see also Borders) inverse projection and, 86–88
conjoint representation and, 25–56 (see language and, 98
also Conjoint representations) location of, 79–86
continuous transition and, 26 meaning of, 80–82
convention and, 147–148, 157 mirrors and, 85
cross-talk and, 71–74 perceived occlusion and, 92
cue integration and, 90 proximal mode and, 84, 93
environmental space and, 67–71, 218–231 reinforcement and, 117
eye and, 79–80 related forms of, 83–86
foreshortening and, 322, 340–345 seeing-in and, 3–15
horizon ratio and, 303–318 simultaneity and, 82
illusion and, 24–25 spatial perception and, 61–75
monocular depth enhancement and, 90 vagueness and, 92
occlusion and, 22–23 window metaphor and, 94–97
parameter setting and, 35 Dürer, Albrecht, 203, 370–371
perspective fallacies and, 198
pictorial relief and, 262–266 (see also Eby, David, 72
Pictorial relief ) Ecological approach, 101, 303
projection and, 207–208 Egyptian method, 375–376
proximal mode and, 56 Eidos, 204, 208
reality and, 27–28 Eisenstein, Sergei, 47
relative, 307–308 Elementary Laws of the Visual Arts
representational primitives and, 36–38 (Cornelius), 358
scale and, 221–222, 236 Emilie Ambre (Manet), 3–4, 13–14
Direct perception, 61–64 Environmental space, 218
Direct realism, 99 borders, 324–326 (see also Borders)
Dissociation, 96–97 information in, 221–225
Distal mode, 85–86, 93 layout perception and, 231–236
Domander, Ramona, 338 lens effects and, 225–229
Drawing, 133–134, 149–150 scale convergence and, 221–222
borders and, 324–336, 346, 348–349, shape of, 219–225
353–354 Environment-centered perception, 67–71
410 Index

Equivalence. See also Dual perception Ghirlandaio, Domenico, 9


concept errors in, 88–89 Gibson, E., 221
cross-talk and, 71–74, 91, 94–96 Gibson, James J., xii–xiii, 61, 237n1, 327,
inverse projection and, 86–88 356
Euclidean space, 261, 262, 278–280, 282 dual perception and, 77
Cayley-Klein geometries and, 283–284 faithful picture and, 127–128
congruences and, 283–287 hybrid theory and, 101–102
Evans, Gareth, 106–107 Marr on, 99
Experience. See Subjectivity optic array and, 135
Experimental phenomenology, 31–32 perspective and, 180–181
Eyes. See also Vision on reality, 218
absolute dimensions and, 308–311 SDT and, 104
Blickfeld and, 356 spatial perception and, 65, 67, 70
as camera, xi–xiii strategic design theory (SDT) and, 103
cross-talk and, 71–74 Gillam, Barbara, 65
curvilinearity and, 181–182 Giorgi, Robert, 309
distance and, 308, 313–314 God, 327
environment-centered perception and, Gombrich, Ernst, 4, 27, 217
67–71 convention and, 105–106, 109
gauge figures and, 252–256 on perspective, 188–189
Goodman on, 182–183 on Thiery, 367–368
hegemony of, 194 Goodman, Nelson, 106, 114
horizon ratio and, 303–318 borders and, 336
inverse projection and, 86–88 on perspective, 158–159, 180, 182–184,
invoked attention and, 114 187–188
iris, xi repleteness and, 174–175
lens, xi symbol theory and, 168
monocular depth enhancement and, 90 Green, Jeffrey, 221, 237n7–237n8
movement and, 182–183 Gris, Juan, 132, 137–138, 141
movies and, 302
multiple pictures and, 287–291 Haber, Ralph, 71
oculomotor adjustment and, 66 Hagen, Margaret, 27, 218, 309
optic array and, 65 natural perspective and, 126, 128–131
perspective theory and, 184–186 (see also Haptic feedback, 62
Perspective) Hauck, Guido, 181, 360–363
pictorial relief and, 291–296 (see also Hecht, Heiko, 405
Pictorial relief ) Hering, Ewald, 39–40
retina, xi, 79–80 Heyer, Dieter, xv, 77–98, 405
rotation around projection center, 363–364 Hildebrand, Adolph, 279–280, 282, 358
seeing modes and, 247–248 Hirstein, William, 113–114
Sehfeld and, 356 Hitchcock, Alfred, 195–196
size and, 307–308, 311–318 Hochberg, Julian, 218
stereoscopic vision and, 63–64 Holbein, Hans, 95
Synopter and, 266 Hopkins, Robert, xvi, 145–165, 405
veridicality and, 248–251 Horizon ratio, 303–306
vista space and, 224–225 absolute dimensions and, 308–311
distance and, 308, 313–314
Farber, James, 30 reality and, 315–318
Figure-ground, 37, 40, 41, 60 size and, 307–308, 311–313
Fiser, Jozsef, 35 How the Women of Dasamonquepeuc Are
Flückiger, Michelangelo, 219–221 Accustomed to Carrying Their Children
Foreshortening, 322, 340–345 (De Bry), 369
Framing, 248. See also Parameters Huffman, David, 137, 139
Husserl, Edmund, 77
Gauge figures, 252–256 Hybrid theory, 99
Geographical space, 119 convention and, 105–118
Gestalt psychology, 31–32, 41, 80 description of, 101–105
411 Index

ecological specificity and, 101 King’s Museum, The (Magritte), 137–138


modularity thesis and, 108–110 Klee, Paul, 131–132, 138–141
nonconceptual spatial perception and, Klein, Felix, 283
110–113 Knowledge. See also Information
occasional, 102 ecological specificity and, 101
motion capture and, 100–101
I funerali dell’anarchico Pinelli (Baj), 50 Koenderink, Jan J., xvii–xviii, 137, 199, 405
Illusions, 24, 77, 83 cross-talk and, 72
common properties and, 172–173 ground theory and, 66
dissociation and, 96–97 on pictorial space, 239–299
resemblance view and, 150–165 Kolyma (Arntz), 44
staring eyes and, 291–296 Kraft, Robert, 221, 237n7–237n8
Thiery on, 367–368 Kratzsch, K., 374
trompe-l’oeil effect and, 5, 82, 96 Kubovy, Michael, 30, 218
Imagery
curvilinearity and, 181–182 La Grotte Chauvet, 232, 234–235
multidimensional, xi Lakoff, George, 333
philosophy of, xii–xiii Lambertian surfaces, 280–281
Imagination, 78, 199–201, 210. See also Landscape with Saint John on Patmos
Perspective (Poussin), 11
IMAX movies, 62 Language, 98, 327
Indirect perception, 61–63 children and, 136–137
Induction, 324–328 meaning and, 41–45
Inferential form, 103 mind’s ear and, 112–113
Information, 302 natural perspective and, 126
borders and, 331–336 (see also Borders) Languages of Art (Goodman), 180
ecological approach and, 303 La Prune (Manet), 11–12
environmental space and, 221–225 Layout, 231–236
horizon ratio and, 303–307 Le blanc-seeing (Magritte), 37–38
pictorial relief and, 241–244 Lens effects, 225–231, 237n8. See also
pictorial space and, 221–225, 241–244, Photography
302–307, 331–336 Leslie, Alan, 45–46
Ingres, Dominique, 11, 13 Levin, Charles, 71
Intention, 49 Levy, Steve, 70
Interior of the Saint Jacob Church in Light, xi
Utrecht (Saenredam), 28–29 brightness, 38–40, 51–56, 135
Intuition, 168–169 chromatic illumination and, 39
Inverse projection, 86–88 contour and, 346–353
Ittelson, William, 218 discontinuities and, 346–348
drawing systems and, 133–134
Jacobs, Robert, 35 eye movement and, 182–183
Jantzen, Hans, 360 modularity thesis and, 108–110
Johnson, Mark, 333 optical edge and, 349–353 (see also
Jones, Rebecca, 27 Borders)
Junctions, 193 optical laws and, 128–143
Jupiter and Thetis (Ingres), 11, 13 perspective fallacies and, 200, 202
Juricevic, Igor, 321–354, 405 pictorial relief and, 267 (see also Pictorial
relief)
Kalkofen, Hermann, xviii, 355–378, 405 picture types and, 244–246
Kandinsky, Wassily, 66 projection and, 207–208
Kant, Immanuel, 327 shadow, 194, 198, 280–281, 346–348
Kapoor, Amish, 83 tone and, 134–135
Karsh, Eileen, 34–35 Linear perspective, xi
Katz, David, 39 curvilinearity and, 181–182
Kemp, Martin, 20 Goodman on, 180, 182–184, 187–188
Kennedy, John M., 218, 321–354, 405 rescue of, 184–189
Kepler, Johannes, xi theory of, 179–181
412 Index

Lines. See Drawing Chatman on, 364–366


Listening Room, The (Magritte), 310 common properties and, 172–173
Loomis, Jack, 71 distal aspects of, 85–86
Lopes, Dominic, 106–107, 111, 121n10 environment-centered perception and,
Lorenz, Konrad, 31 67–71
Lyons, John, 364–366 floating in space, 65–67
language meaning and, 41–45
Mach, Ernst, 355–357 line junctions and, 337
Mach card, 39–40 mirrors and, 85
Mackavey, William, 358 multiple moments and, 368–375
Magritte, René, 37–38, 132, 137–138, 310 opaque, 240
Manet, Édouard, 3–4, 11–14 optical laws vs. symbolic rules, 125–
Mantegna, Andrea, 365 143
Marey, Etienne Jule, 365 proximal mode and, 85–86
Markel, Angela, 371 repleteness and, 174–175
Marr, David, 28, 99, 121n10 representational primitives and, 40–41
on art, 133 resemblance view and, 150–165
strategic design theory (SDT) and, 103–104 as signs, 170–175
viewer-centered representation and, 64, 70 spatial perception and, 65–75
Mausfeld, Rainer, xiv, 17–60, 406 stereopsis and, 66
Maynard, Patrick, xvii, 191–211, 406 superiority effects and, 40
Measurement theory, 215–216 as symbol, 170
absolute dimensions and, 308–314 Occasional constructivism, 102
differentiation and, 257–259 Occlusion, 22
environmental space and, 219–221 borders and, 330 (see also Borders)
gauge figures and, 252–256 dimension and, 37–38
meaning of, 261–262 perceived, 92
pictorial relief and, 262–266 (see also subjectivity and, 202
Pictorial relief ) Oh! but oh! (Klee), 138
relative distance and, 308 Opaqueness, 240
relative size and, 307–308 Optic array, 65
scale convergence and, 221–222 absolute dimensions and, 308–311
Meltzoff, Anchew, 47, 49 distance and, 308, 313–314
Meng, Jeanette, 67 size and, 307–308, 311–313
Metarepresentation, 46 Oursler, Tony, 97
Michotte, Albert, 20–21, 31 Outlines. See Borders
Milbrath, Constance, 346
Mirrors, 85 Palmer, Stephen, 89
Mit Grunen Strümpfen (Klee), 140 Parameters
Modularity thesis, 108–110 color and, 38–40
Moskowitz, Breyne, 136–137 cross-talk and, 71–74, 91, 94–96
Motion capture, 100–101 dimension and, 37–38
Motor skills, 32–36 dissociation and, 96–97
Mourning of Dead Christ (Mantegna), 365 figure-ground ambiguity and, 41
objects and, 40–41
Nakayama, Ken, 36 redeployment and, 115–116
Naked on the Bed (Klee), 138 setting of, 30–36
Natural perspective, 126, 128–131 surfaces and, 40–41
Necker cube, 93 vagueness and, 51–56
Newman, Barnett, 8 Pavimenti, 191
Nicholls, Andrea, 71–73, 341–342 Peacocke, Christoper, 106–107
Niederée, Reinhard, xv, 77–98, 406 Pearson, Don, 135
Peephole, 194, 204
Obermair, Christoph, 338 Perception, xi, 57–60
Objects. See also Pictorial relief; Surfaces absolute dimensions and, 308–314
borders and, 331–336 (see also Borders) borders and, 324–326 (see also Borders)
413 Index

cognitive architecture and, 36–51 convention and, 105–118


commonsense taxonomies and, 17–19 hybrid theory and, 101–105
conjoint representations and, 25–30 (see modularity thesis and, 108–110
also Conjoint representations) motion capture and, 100–101
content strategies and, 99–122 nonconceptual spatial perception and,
control mechanisms and, 116 110–113
convention and, 105–118 pictorial space and, 105–119
cross-talk and, 71–74, 91, 94–96 redeployment and, 115–116
current model of, 79–80 Perner, Josef, 335
differentiation and, 257–259 Perspective, xi, xiii
direct, 61–64 absolute, 362–363
dissociation and, 96–97 alternative views on, 207–210
dual nature of, xiv–xv (see also Dual conjoint representations and, 25–30
perception) convention and, 145–165
ecological approach to, 101, 303 curvilinearity and, 181–182
environmental space and, 67–71, 218– deformations and, 194
231 drawing systems and, 133–134
figure-ground ambiguity and, 41 dual perception and, 89–92 (see also Dual
horizon ratio and, 303–318 perception)
hybrid accounts and, 101–122 experience and, 160–161
imagination and, 78, 199–201, 210 eye movement and, 182–183
imputed mental states and, 46–51 faithful, 127–128
indirect, 61–63 fallacies of, 195–207
intention and, 49 Gombrich on, 188–189
inverse projection and, 86–88 Goodman on, 180, 182–184, 187–188
language/meaning and, 41–45 Hauck on, 360–363
measurement theory and, 215–216 heuristics of, 195–203, 207–210
multiple pictures and, 287–291 horizon ratio and, 303–318
nonconceptual, 110–113 intuition and, 168–169
parameter setting and, 30–36 linear, 179–189
photographic space and, 218–231 mental states and, 46–51
pictorial relief and, 259–260 (see also multiple, 358
Pictorial relief ) natural, 126, 128–131
pretense play and, 45–46 oblique projection and, 131–132
proximal mode and, 54–56, 79–80, 85–86, optical laws vs. symbolic rules, 125–143
93 optic array and, 135
psychophysics and, 37–41 real uses of, 192–193
reality and, 27–28, 216–218 Renaissance and, 19–20, 179–180
redeployment and, 115–116 resemblance view and, 150–165, 167–
reinforcement and, 117 177
relative distance and, 308 robustness, 94, 95, 97
relative size and, 307–308 seeing-in and, 3–15, 147
representational content and, 3–13 spectator’s education and, 157–158,
representational primitives and, 32–36 193–195
scale convergence and, 221–222 status of, xv–xvii
seeing-in and, 3–15, 147 strategic design theory (SDT) and,
spatial layout and, 61–75 (see also Spatial 119–120
perception) subjectivity and, 202, 360–362 (see also
surfaces and, 14–15 Subjectivity)
triggering and, 30–36 technology and, 191–193, 206, 210
twofoldness and, 3–4 theory of, 146–149, 184–186
vagueness and, 51–56 vagueness and, 193–194
visibility limits and, 9–13 validity and, 145
Perception of the Visual World (Gibson), 356 variety accommodation and, 149–156
Perceptual strategies Perspective, New Principles of Linear
control mechanisms and, 116 Perspective (Taylor), 197–198
414 Index

Philbeck, John, 71 dual character of, 19–21 (see also Dual


Philosophy, xii, 215 perception)
Cornelius and, 357–358, 360 ecological specificity and, 101
Hauck and, 360–363 Egyptian method and, 375–376
hybrid accounts and, 99, 101–122 experience and, 160–161
Kant and, 327 exploration of, xvii–xviii
Mach and, 355–358 eye rotation and, 363–364
Pirenne and, 363–364 faithful representation and, 127–128
Photography, xi–xiii, 216–217, 301 figure-ground ambiguity and, 41
camera obscura and, 126–127 future research for, 296–299
distortion and, 310 Hauck and, 360–363
environmental space and, 218–231 Hildebrand and, 279
horizon ratio and, 303–318 horizon ratio and, 303–318
lens effects and, 225–231, 237n8 hybrid theory and, 101–119
perceived spatial shape of, 229–231 ideography and, 366–367
tone and, 134–135 illusion and, 24
Physicalistic trap, 31 imputed mental states and, 46–51
Picasso, Pablo, 48 information in, 221–225, 241–244,
Pictorial relief, 239–240 302–307, 331–336
beholder’s share and, 277–278 meaning in, 41–45, 318–320
cognition vs. perception, 259–260 modularity thesis and, 108–110
cueing and, 251, 266–267, 280–281 motion capture and, 100–101
differentiation orders and, 257–259 multidimensional, 22–23 (see also
global properties and, 256–257, 267–272 Dimension)
Hildebrand and, 280, 282 multiple moments and, 368–375
information and, 241–244 multiple pictures and, 287–291
local properties and, 256–257, 267–272 natural theories of, 126, 128–131
measurement theory and, 261–262 nonconceptual spatial perception and,
operational methods for, 251–278 110–113
picture types and, 244–247 oblique projection and, 131–132
seeing models and, 247–248 occlusion and, 22, 37–38, 92, 202, 330
staring eyes and, 291–296 optical laws vs. symbolic rules, 125–143
task variation in, 272–277 perceived depth and, 20–21
veridicality and, 248–251 performance dimension and, 114
viewing modes and, 264–266 perspective and, 179–189 (see also
Pictorial space, xi, 377–378 Perspective)
absolute dimensions and, 308–314 philosophy of, xii–xiii
alternative views of, 207–210 photography and, 218–231 (see also
art and, xii (see also Art) Photography)
attentional shift and, 81 Pirenne and, 363–364
borders and, 324–326 (see also Borders) planar aspects of, 81–82
Cayley-Klein geometries and, 283–284 reality and, 27–28
Chatman and, 364–366 redeployment and, 115
close distance constructions and, 357–360 reference and, 175–177
composition and, 191 relative distance and, 308
concept of, xii–xiii relative size and, 307–308
congruence groups and, 279–287 religion and, 217
conjoint representations and, 25–56 (see representational content and, 3–13,
also Conjoint representations) 332–36
content and, 113–119, 175–177 resemblance view and, 150–165, 167–177
control mechanisms and, 116 scenography and, 364–366
convention and, 105–113 seeing-in and, 3–15, 147
Cornelius and, 357–360 signs and, 170–171, 173–175
cross-talk and, 71–74, 91, 94–96 spatial layout and, 61–75
cue integration and, 21–25 spectator’s education and, 157–158,
curvilinearity and, 181–182 193–195
415 Index

status of, xv–xvii Rehkämper, Klaus, xvi–xvii, 179–189,


structure of, 278–291 406
subjectivity and, 202 (see also Subjectivity) Reinforcement, 117
surfaces and, 321–354 (see also Surfaces) Relativism, 204
as symbol, 170 Religion, 217
tallying effect of, 8 Renaissance, 77
Thiery and, 367–368 curvilinearity and, 181
trompe-l’oeil pictures and, 5 inverse projection and, 87
unrepresented content and, 118–119 perspective and, 19–20, 179–180
variety accommodation and, 149–156 Repleteness, 174–175
visual fields and, 355–357 Representational content
Pirenne, Maurice, 30, 77, 210 art and, 3–13
on eye rotation, 363–364 conjoint representations and, 25–30
photography and, 358–359 surfaces and, 14
Renaissance perspective and, 180 tallying effect and, 8
Polanyi, Michael, 45, 77 visibility limits and, 9–13
Portrait from within, 355–358 Representational primitives
Portrait of an old man and his grandson color and, 36, 38–40
(Ghirlandaio), 9 dimension and, 37–38
Postmodernism, 194–195 imprecision and, 151–156
Potemkin (Eisenstein), 47 imputed mental states and, 46–51
Poussin, Nicolas, 11 language and, 41–45
Pozzo, Andrea, 82, 224 modularity thesis and, 108–110
Prazdny, Kvetoslav, 23 nonconceptual spatial perception and,
Pretense play, 45–46 110–113
Prisms, 368 objects and, 40–41
Projection, 194, 195f, 206 optical laws vs. symbolic rules, 125–143
Proust, Marcel, 6 parameter setting and, 32–36
Proximal mode, 79–80 pretense play and, 45–46
conjoint representations and, 54–56 resemblance view and, 150–165
distal mode and, 93 surfaces and, 40–41
dual perception and, 85–86 vagueness and, 51–56
Psychology, xii, 215. See also Cognition; Resemblance view, 150, 161–165, 167
Gibson, James J.; Pictorial relief content and, 175–177
figure-ground and, 41 convention compromise and, 156–160
Gestalt, 31–32, 41, 80 imprecision and, 151–156
photographic space and, 216–231 intuition and, 168–169
seeing-in and, 6, 8 picture theory and, 169–175
Psychophysics reference and, 175–177
color and brightness, 38–40 Reunion of foreign birds placed within
cue integration and, 21–25 different cases (de Barde), 7
depth, 37–38 Reverse angle shots, 369–370
objects, 40–41 Rogers, Sheena, xviii, 218, 301–320, 406
surfaces, 40–41 Rollins, Mark, xv, 99–122, 406
Rosinski, Richard, 30
Racinet, A., 373
Ramachandran, Vilayannr S., 36, 103–104, Sachs-Hombach, Klaus, xvi, 167–177,
113–114, 117 406
Raphael, 159–160 Saenredam, Pieter, 28–29
Realism, 99 Sartre, Jean-Paul, 78
Reality Scenography, 364–366
horizon ratio and, 303–318 Scheiner, Christoph, 360
size/distance perception and, 307–308, Schier, Flint, 106–107, 111
314–318 School of Athens, The (Raphael), 159–160
Redeployment, 115–116 Schriever, Walter, 23
Reference, 175–177 Schwartz, Robert, 406
416 Index

Schwitters, Kurt, 206 hybrid theory and, 101–119


Sedgwick, Hal, xiv–xv, 61–75, 406 inferential form and, 103
Seeing-in, 3, 147 invoked attention and, 114
artist’s intent and, 6, 8 Marr on, 103–104
dual approach to, 5–6 modularity thesis and, 108–110
representational content and, 5–13 motion capture and, 100–101
tallying effect of, 8 nonconceptual spatial perception and,
twofoldness and, 4–15 110–113
Sehfeld, 356 performance dimension and, 114
Sejnowski, Terrence, 112–114 pictorial space and, 105–119
Serveto and His Execution (van Sichem), Ramachandran on, 103–104
374 redeployment and, 115–116
Shading, 108–110 reinforcement and, 117
Shadow. See also Borders unrepresented content and, 118–119
discontinuities and, 346–348 Subjectivity. See also Views
perspectival deformations and, 194 beholder’s share and, 277–278
perspective fallacies and, 198 Hauck on, 360–362
pictorial relief and, 280–281 (see also multiple pictures and, 287–291
Pictorial relief ) observer goal and, 259–260
Shape. See also Spatial perception pictorial relief and, 251–256 (see also
borders and, 324–326 (see also Borders) Pictorial relief )
environmental space and, 219–225 picture types and, 244–246
optical laws vs. symbolic rules, 125–143 spectator education and, 157–158,
photographic space and, 229–231 193–195
scale convergence and, 221–222 Summer (Arcimboldo), 98
Shimojo, Shinsnke, 36 Surfaces. See also Pictorial relief
Signs, 170–175 borders and, 321–336, 346–349, 353–354
Smith, Robert, 365 contour ambiguity and, 349–353
Spatial perception, 125–127. See also curved lines and, 336–337
Pictorial relief; Pictorial space foreshortening and, 322
action space and, 224 imputed mental states and, 46–51
cross-talk and, 71–74 induction and, 324–327
dissociation and, 96–97 Lambertian, 280–281
environmental space and, 67–71, 219–231 line junctions and, 337
floating objects and, 65–67 optical edge and, 349–353
foundational theories for, 65–67 outline and, 324–326
multiple moments and, 368–375 perspective fallacies and, 196–197
nonconceptual, 110–113 representational primitives and, 40–41
oblique projection and, 131–132 simultaneous location and, 40
optical laws vs. symbolic rules, 128–143 straight lines and, 336–337
planar aspects of, 81–82 texture gradients and, 331–332
resemblance view and, 150–165 Swedlund, Charles, 226–227
vista space and, 224–225 Synopter, 266
Spatial unity, 120n8
Stemberger, G., 372 Talbot, William Henry Fox, 211n6
Stereopsis, 23, 24, 25, 66 Tallying effect, 8
Stern, William, 136 Taylor, Brook, 127, 197–200
Stevens, Kent, 23–24 Technology, 249–250
Stevens, Stanley, 215–216 perspective and, 191–193, 206, 210
Stimuli. See Perception television, 193, 197
Story and Discourse (Chatman), 364 Ten Doesschate, Gezienus, 355
Strategic design theory (SDT ), xv, 122n14 Texture, 331–332
control mechanisms and, 116 Thiery, Armand, 367–368
convention and, 105–118 Tinbergen, Nikolaas, 31
direct component of, 102–103 T-junction, 139–141
ecological specificity and, 101 Topology, 283–287
417 Index

Transmission of the Divine Spirit, The borders and, 328–336 (see also Borders)
(Pozzo), 82 computer experiments and, 71–72
Transparency, 40 conjoint representations and, 25–56 (see
Triggering, 30–36, 51 also Conjoint representations)
Trompe-l’oeil pictures, 5, 58, 82, 96, 172 convention and, 105–113
Turhan, Mümtaz, 23 cross-talk and, 71–74, 91, 94–96
Twofoldness, 3 directed, 104
seeing-in and, 4–15 ecological specificity and, 101
tallying effect and, 8 environment-centered perception and,
Two Venetian Women (Dürer), 370 67–71
figure-ground ambiguity and, 41
Ultramodernism, 195 illusion and, 24
imagination and, 78
Vagueness, 51–53 induction and, 324–328
perceived, 92 mirrors and, 85
perspective and, 193–194 monocular, 24, 328
proximal mode and, 54–56 motion capture and, 100–101
van Doorn, Andrea J., xvii–xviii, 66, occlusion and, 22, 37–38, 92, 202, 330
239–299, 405 optic array and, 65
Van Essen, David, 113, 115 parameter setting and, 30–36
Vanishing point. See Perspective perspective theory and, 184–186 (see also
van Sichem, Christof, 374 Perspective)
Veridicality, 26, 57, 248–251 psychophysics and, 37–41
Views, 64, 70, 377–378 redeployment and, 115–116
alternative approaches and, 207–210 representational content and, 3–13
beholder’s share and, 277–278 retinal patterns and, xi
close distance constructions and, 357–360 spatial layout and, 61–75
Egyptian, 375–376 stereoscopic, 63–64, 66
gauge figures and, 252–256 vagueness and, 51–56, 92
Hauck and, 360–363 visibility limits and, 9–13
horizon ratio and, 303–318 Vista space, 224–225
ideography and, 366–367 Visual fields, 355–357
modes of, 264–266 Visual rays, 281
multiple moments and, 368–375 Visual semiworlds, 99–100
multiple pictures and, 287–291 von Brentano, Franz, 243
optical array and, 307–314 von Helmholtz, Hermann, xi, 39, 221, 356
perspective fallacies and, 203–207 (see also von Uexküll, Jakob, 31
Perspective)
pictorial relief and, 264–266 (see also Wallach, Hans, 34–35
Pictorial relief ) Walton, Kendall, 301, 319
POV shots and, 369–370 Weeping Woman (Picasso), 48
resemblance, 150–165, 167–177 Wickhoff, Franz, 374–375
reverse angle shots and, 369–370 Willats, John, xvi, 125–143, 218, 406
rotation and, 363–364 Wimmer, Heinz, 335
scenography and, 364–366 Window metaphor, 94. See also Parameters
seeing modes and, 247–248 cross-talk and, 71–74, 91, 94–96
theory of pictures and, 146–149 dissociation and, 96–97
Thiery and, 367–368 spatial dissociation and, 96–97
veridicality and, 248–251 With Green Stockings (Klee), 138–139, 141
visual fields and, 355–357 Wollheim, Richard, xiv, 3–15, 147, 406
zoography and, 364–366
Vir Heroicus Sublimis ( Newman), 8 Yellott, John, 23
Virtual reality, 63
Vishton, Peter, 221 Zöllner illusions, 367
Vision. See also Dual perception; Eyes Zoography, 364–366
binocular, 329

You might also like