221 85 59MB
English Pages 602 [604] Year 2008
Instruments in Art and Science
≥
Theatrum Scientiarum English Edition Edited by Helmar Schramm, Ludger Schwarte, Jan Lazardzig Scientific Advisory Board Hartmut Böhme, Olaf Breidbach, Georges Didi-Huberman, Peter Galison, Hans-Jörg Rheinberger, Wilhelm Schmidt-Biggemann, and Barbara Maria Stafford
Volume 2
Walter de Gruyter · Berlin · New York
Instruments in Art and Science On the Architectonics of Cultural Boundaries in the 17th Century Edited by Helmar Schramm, Ludger Schwarte, Jan Lazardzig
Walter de Gruyter · Berlin · New York
앝 Printed on acid-free paper which falls within the guidelines 앪 of the ANSI to ensure permanence and durability.
Library of Congress Cataloging-in-Publication Data [Instrumente in Kunst und Wissenschaft. English] Instruments in art and science : on the architectonics of cultural boundaries in the 17th century / edited by Helmar Schramm, Ludger Schwarte, Jan Lazardzig. p. cm. ⫺ (Theatrum scientiarum ; v. 2) Includes bibliographical references and index. ISBN 978-3-11-020240-3 (alk. paper) 1. Scientific apparatus and instruments ⫺ History ⫺ 17th century. 2. Art and science ⫺ History ⫺ 17th century. I. Schramm, Helmar. II. Schwarte, Ludger. III. Lazardzig, Jan. IV. Title. Q185.I59 2008 681.091032⫺dc22 2007049145
ISBN 978-3-11-020240-3 Bibliographic information published by the Deutsche Nationalbibliothek The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data are available in the Internet at http://dnb.d-nb.de. 쑔 Copyright 2008 by Walter de Gruyter GmbH & Co. KG, D-10785 Berlin All rights reserved, including those of translation into foreign languages. No part of this book may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording or any information storage and retrieval system, without permission in writing from the publisher. Printed in Germany Cover design: Christopher Schneider, Berlin Printing and binding: Hubert & Co., Göttingen
Editors’ Preface This volume constitutes the second part of an intended total of eight in the series Theatrum Scientiarum, which sets out to examine the fundamental crossover of art and science in a new way. The project is based on the assumption that in the course of the reconstitution of science in the seventeenth century practices of presentation, observation, and medial competence emerge, whose productive force can only be adequately described from an interdisciplinary perspective. These practices are in no way confined to the processes of the legitimization and implementation of knowledge; rather, through the experimental methods used for modeling and handling the world, a dynamic structure of creative approaches to observation and presentation is developed. The questions we would like to address in the series Theatrum Scientiarum emerge from the cultural upheavals of our times. They are carried by the conviction that an appropriate understanding of the interaction of today’s medial configurations of scientific programs and artistic practice is only possible in the awareness of this long-term historical process. It is due to the international character of the questions discussed that this volume, which appeared in German in 2006 after a conference in Berlin, now also appears in English. The positive reception given to the first English-language version in the spring of 2005 with the title “Collection, Laboratory, Theater,” gives us reason to hope that this volume will also contribute to the discussion on the complex relationship between art and science. Therefore we are highly indebted to the chief editor of literary studies of the Walter de Gruyter publishing house, Dr. Heiko Hartmann, for his capable and trusting cooperation as well as for his interest in the project. The realization of this long-term research project would not have been possible without the generous support of the Freie Universität Berlin, the German Research Foundation (DFG), as well as the Fritz Thyssen Foundation. The Musikinstrumenten-Museum Berlin SIMPK, namely Professor Conny Restle, with the loan of the Curt Sachs lecture hall made an essential contribution to the success of the conference. The Deutsches Technikmuseum Berlin, particularly Dr. Maria Borgmann,
VI
Preface
must be thanked that the conference found such a stimulating setting – and conclusion – in the newly fitted exhibition area. We would like to thank the translators of this volume (they are named at the end of each contribution) for their outstanding work. The editors warmly thank Daniel Hendrickson and Bill McCann for conscientiously taking care of the amalgamation and supervision of the translations. And as always we are highly grateful to Michael Lorber and Daniela Hahn for their excellent editorial supervision of this volume, which made this publication possible in such a short time. With the English publication, we would like to make a pragmatic plea for the plurality of scientific languages. The clear preponderance of German literature in the notes repeatedly presented the translators with difficulties. When no appropriate English translation was available, the quotations were, without renewed quotation from German (or Italian, French, Spanish, etc.) translated from the original. Only with quotations from unpublished or rare sources (e.g. archives) is, as a rule, the original quotation also recorded in the footnotes. The Editors
Contents
Editors’ Preface ..........................................................................................
V
Contents ...................................................................................................... VII Helmar Schramm Introduction: The Hand as “instrumentum instrumentorum” .....................
XI
Hans-Jörg Rheinberger Intersections: Some Thoughts on Instruments and Objects in the Experimental Context of the Life Sciences ................................................
1
Dieter Mersch Representation and Distortion: On the Construction of Rationality and Irrationality in Early Modern Modes of Representation ......................
20
Olaf Breidbach World Orders and Corporal Worlds: Robert Fludd’s Tableau of Knowing and its Representation .................................................................
38
Florian Nelle Telescope, Theater, and the Instrumental Revelation of New Worlds ...........................................................................................
62
Frank Fehrenbach The Pathos of Function: Leonardo’s Technical Drawings .........................
78
Nicola Suthor “Il pennello artificioso”: On the Intelligence of the Brushstroke ............... 106 Barbara Maria Stafford The Enlightenment “Catholization” of Projective Technology: Theurgy and the Media Origins of Art ....................................................... 127 Jan Lazardzig The Machine as Spectacle: Function and Admiration in SeventeenthCentury Perspectives on Machines ............................................................. 152
VIII
Contents
Ludger Schwarte The Anatomy of the Brain as Instrumentalization of Reason .................... 176 Gerald Hartung The “Chymistry Laboratory”: On the Function of the Experiment in Seventeenth-Century Scientific Discourse ................................................. 201 Gerhard Wiesenfeldt The Order of Knowledge, of Instruments, and of Leiden University, ca. 1700 ....................................................................................................... 222 Angela Mayer-Deutsch The Ideal Musaeum Kircherianum and the Ignatian Exercitia spiritualia .................................................................................... 235 Conny Restle Organology: The Study of Musical Instruments in the 17th Century ......... 257 Andreas Meyer In Sound Similar to the Harps: Early Descriptions of African Musical Instruments ................................................................................... 269 H. Otto Sibum Machines, Bats, and Scholars: Experimental Knowledge in the Late Eighteenth and Nineteenth Centuries ......................................................... 280 Peter Galison/Lorraine Daston Scientific Coordination as Ethos and Epistemology .................................. 296 Stefan Ditzen Breaking, Grinding, Burning: Instrumental Aspects in Early Microscopical Pictures ............................................................................... 334 Jochen Hennig The Instrument in the Image: Revealing and Concealing the Condition of the Probing Tip in Scanning Tunneling Microscopic Image Design ..... 348 Bruno Bachimont Formal Signs and Numerical Computation: Between Intuitionism and Formalism. Critique of Computational Reason ................................... 362 Don Ihde Art Precedes Science: or Did the Camera Obscura Invent Modern Science? ...................................................................................................... 383 Thomas F. Gieryn Instrumentalities of Place in Science and Art ............................................. 394
Contents
IX
Georges Didi-Huberman The Eye Opens, the Lamp Goes Out: Remarks on Bergson and Cinematography ......................................................................................... 421 Martin Burckhardt The Illusion of Power: Central Bank Money ............................................. 437 Sybille Krämer The Productivity of Blanks: On the Mathematical Zero and the Vanishing Point in Central Perspective. Remarks on the Convergences between Science and Art in the Early Modern Period ............................... 457 Jörg Jochen Berns Instrumental Sound and Ruling Spaces of Resonance in the Early Modern Period: On the Acoustic Setting of the Princely potestas Claims within a Ceremonial Frame ............................................................ 479 About the Authors ...................................................................................... 507 Image Credits .............................................................................................. 515 Bibliography ............................................................................................... 517 Index of Names ........................................................................................... 555 Index of Subjects ........................................................................................ 563
HELMAR SCHRAMM
Introduction: The Hand as “instrumentum instrumentorum” I. The trace of a human hand. Plaster casts of human hands. In the middle of the nineteenth century Carl Gustav Carus collected a whole series of such casts. He arranged them strictly according to a scale of values, to direct his inquiry to an original point of intersection where a typology of hands overlaps with the typical use of instruments, tools, things, and for each the habitual way in which they are handled. The concrete order of grips could then be interpreted as a catalogue of exemplary social types, affording insights into society’s different modes of functioning. The graphic patterns of the collection extended from the delicate, slightly nervous hand – almost reaching for the piano by itself – to the coarse hand of the worker, callused from handling heavy tools. Now let us confront this cultural-historical gallery of hands with other, more recent hand sketches to find additional pointers for our theme. Firstly, the picture of an almost naked hand, caught just at the point of being uncovered, pulling off an elegant glove. An image condensed by Pierre Klossowski into a living symbol; an image from which a broadranging presentiment of the erotic emerges, and where every possibility of sensual presence and human productivity is sublimated into a single, silent gesture. Juxtaposed with this – and in sharp contrast – imagine a hand bared of all covering, and, in addition, all disturbing corporeality: a hand as skeleton, the reliable gripping instrument of a robot, representing consummate functionality. The precise mechanics and multiple nature of the metallic robot hand speak of a functional dynamics, but also forcefully recall the coldness of death. Finally, next to this symbol of mechanical rationality, a glove-covered hand again, but this time with the status of violent omnipotence, a divine hand beneath a data-glove. The hand as symptomatic ‘instru-
XII
Helmar Schramm
mentum instrumentorum’ of our time, suggesting that nothing is beyond its power, that our whole being can literally be focussed on the presence of this hand, this glove. Such ambivalent images raise questions that undoubtedly affect the whole of today’s culture, its architectonics, boundaries, and possibilities. The complexity of these problem zones, however, is bound up with questions that cannot be answered from the perspective of the present alone. It is therefore valid to consider cultural-historical deep spaces, without however losing sight of the highly charged dynamics of contemporary developments. Here, since the beginning of the modern period in Europe, points of reference and question marks, affording a certain distance, can be found in numerous machine and firework books, corresponding in manifold ways to the aforementioned symbol of a sophisticated, instrumentalized data-hand. So, for example, in a book on fireworks and weapons technology from 1602, the talk, expressly following Aristotle, is of the hand as ‘instrumentum instrumentorum,’ and, in the context of our publication on the role of scientific and artistic instruments in the historic formation of an architectonics of cultural boundaries, this deserves special attention, as here, starting with the hand as an instrument of creative violence, the whole spectrum of instruments at the beginning of the seventeenth century is inventoried. If such an overview of instruments happens to appear in a work on weapons technology and artificial fireworks, this need not surprise us, since even then war was a fundamental event where local peculiarities were condensed to the sum of a generalizable experience. All dimensions of life, all the concretizations of human activity were interwoven in one way or another with the logistical and technical implications of war. A war instrument was thus anything related to architecture, fortification, or the whole concrete arsenal of tools necessary under siege conditions to maintain everyday life in terms of survival; meaning that literally anything that was perceived by a community as an instrument at all could be interpreted as a war instrument. However, what is particularly interesting in relation to the aforementioned early typology is not only the tendency towards completeness, the wish for a total overview as expressed by the plethora of numerous comparable illustrations and explanations. Interest is due particularly to the anthropomorphic gesture affecting the world of instruments as a formative power. Here it was visibly the case of something resembling the unfurling of a corporeal presence, of the projection of developed human sensuality onto the surface of a system of instruments, initially
Introduction
XIII
considered purely typologically. Tending in the same direction is the fact that the presentation of individual instruments repeatedly found their starting point in the exposition of practical operations or imaginable operative possibilities. If one follows these traces of connections (here merely hinted at) between action and instrument more attentively, then one almost inevitably comes up against a strange paradox in the reception of Aristotle that must have been of considerable significance for the concrete developments of experimental science and art in the seventeenth century. On the one hand, one could say that Aristotle’s natural-philosophical approach gradually, little by little, suffered defeat, to be replaced by a new thinking that in an almost exemplary way inspired Galileo’s famous dialogue on the two world systems. On the other hand, in the seventeenth century we find the starting point for a far-reaching career of aesthetic elements of the Aristotelian concept. Combined with the feverish attacks of a passionate, omnipresent theatrical culture, and the search for effective antidotes for the treatment of ambivalent passions, the Aristotelian conception of catharsis and tragedy was received under new auspices. With a closer inspection of the relevant poetics of the time, it becomes clearer what is at issue here, namely the preparation and theatrical testing of instrumentalizable concepts of action. If Aristotelian poetics are concerned with the imitation of great actions, then the question remains of what this imitation should actually look like, and how, as it were, in the spirit of a theatrum anatomicum, it should be dissected, so as to be controllably performed. And Aristotle expressly emphasized that in the case of tragedy, the appropriate imitation should not arise through report, but through the acting characters themselves. All this undoubtedly casts light on the relation between hand and word that marks culture. I do not want to go into the details at this point, but rather, from the current perspective regarding a global endgame of Western culture, open up for discussion the thesis that in the seventeenth century a broad-ranging interplay of instrument and action unfolded, and this under the auspices of tragedy. II. The trajectory of this tendency becomes foreseeable when one compares the aforementioned instrument book from the dawn of the seventeenth century to another large project, Jacob Leupold’s Theatrum machinarum, where a totally new situation is reflected, which, from the
XIV
Helmar Schramm
end of the century onwards, became increasing valid. In Leupold’s manyvolumed work, the initially declared intention is to compile and itemize all the faculties of the mechanical arts systematically to make them available for good causes. Nevertheless a broad-ranging shift in accent becomes apparent, intimated by an inconspicuous symptom: the physical hand as an anthropomorphic center disappears in the new cosmos of instruments and machines. The industrious hand is dissolved into a superior principle of objectivization and concretization, and thus no longer appears in essence. This path also has to do with the way certain factors of the hand’s movement are recognized, inventoried, and rationalized as mere grips, as highly specialized functions. These volumes are doubtless a symptom of the growing power of mechanization, but also the drawing up of a complete fragmentation of action and patterns of behavior, which should be taken into account when thinking about instruments and the possibilities and limits of instruments. In the seventeenth century drastic changes took place in the field of the experimental sciences, and the founding of experimental science arose out of the spirit of mathematics and mechanics. A new order of instrumental laboratory setup emerges, which was first of all the invention of the laboratory in an interplay with the development of new instruments. Apart from devices of measuring, weighing, calculating, and recording, other elaborate devices appear, such as the telescope, microscope, and air pump, which have a lasting influence on the relations between observer, mode of perception and object of research. Against the background of an increasingly thorough systematization of experimental strategies, instruments are seen as a predictable, reliable means to achieve results in an optimal way. This calculable, functional aspect could also be transferred to areas of essential significance for concrete human life-situations, for example surgical practice or the legal system. Here a rationalizing trend appears that also problematizes the relationship to culture-forming action processes such as all-prevalent rituals. If the laboratory stands for a release from the magic-ritual relationship to nature represented by the magia naturalis, then the spheres of society, politics, and political philosophy are concerned with releasing ritual potential from its religious connections. Due to a strong awareness of the enormous ritual practices have on all human senses, centuries of experience had shown how ecclesiastical space in the truest sense of the word was cultivated as the architectural pendant to a developed human sensuality. This tendency towards transforming religious rituals can be described as an attempt to change ritual into a practicable rhetorical in-
Introduction
XV
strument in all its parts, in view of performative techniques, visual strategies, and the normative aspects of language. Precise considerations on this subject can be found in Thomas Hobbes, which more or less boil down to summarizing officials as ‘fictitious characters,’ and contracts and institutions as complex ritualized operations. Interestingly, after 1700, precisely this approach is passed through a filter of educational concepts and ideals, and systematically developed into the foundation of various types of Enlightenment theater, which, at heart, are nothing other than complex rhetorical instruments whose effective construction extends from stage-architectural principles to acting techniques. With the idea of the theater as an instrumentalized cultural asset, the hand turns up again, though now in a different form. If one considers the great acting theories produced between 1750 and 1800 for instance, the hand emerges as if bound up in the most varied models of gestural tables and arrangements of signs. Here acting is treated as something exclusively symbolic. The idea of a symbolic action hangs on this instrumentalized hand, which is now instrumentalizable in a new way because at heart it is increasingly only an idea, its physical presence is a symbol. The educational ideal articulated in this way and placed on the theatrical stage now collides with a totally different ‘theater’ whose basic structures appear in Leupold’s previously mentioned Theatrum machinarum. An essential aspect of the change of the instrument in the seventeenth century consisted in the fact that – on the stable foundations of mathematical calculations and a new understanding of mechanical relationships – instruments and machines gradually found mass circulation beyond the walls of the laboratory into all areas of culture, producing specific dynamics. Leupold’s Theatrum machinarum shows in what sense, as early as the turn of the eighteenth century, mechanical engineering is applied, whose broad-ranging expansion into a completely changed ‘rhythmicization’ of the whole of life-culture is still visible one hundred years later. For people in their corporeal existence, society as a universe of cultural rhythms also means integration into contradictory, mutually conflicting rhythmic systems and related dissonances, dissections, fragmentations. Every culture is provided with a specific instrument, a collection of instruments, which play an essential role in the shaping of cultural rhythms. One naturally thinks, first of all, of musical instruments, but the real scope of the problem is only treated properly when a culture’s musical instruments are placed within the whole complexity of the respective theatrum machinarum. In this sense, and in reference to the in-
XVI
Helmar Schramm
strumentality developed in the seventeenth century, questions about processes of ritualization can also be raised that might contribute to the deepening of our understanding of contemporary cultural rhythms, their tempo, their speed, their obsolescence. III. Jörg Jochen Berns highlights interactions between instrument and ritual by revealing the outlines of an acoustic doctrine of instruments from courtly ceremonial texts and festivity diaries. What is interesting here is his comparative reflection on musical instruments and other preparations, apparatuses, and aids to sound production. What emerges from this is references to the musicological and physical discourses of the time, even if these are not elaborated in the contribution itself. Instead the accent is placed on the behavior-forming effects of a sophisticated instrument of acoustic sign-production. The main emphases of the investigation focus on scenic sound-situations, typical echo-spaces, and finally an acoustic doctrine of instruments. The field of investigation affords, in addition, a consideration of relations between instrumental techniques and dispositives of power. Precisely this leads to the core-questions constantly bound up with the deployment of instruments. In this study, different instruments are brought into play, to whose heterogeneity an excursus on the precision of the instrument-term responds. Here it is underlined “that the differences between terms such as tool, instrument, implement, machine, apparatus, device in everyday practice, but also in scientific practice, are imprecise” (501). The author emphasizes the productivity of this imprecision, though he also notes how over centuries musical instruments were terminologically grasped quite clearly, which may have resulted in a special status in the contention about the definition of the term ‘instrument.’ An instructive overview of organology – musical instrument lore – in the seventeenth century is provided by Conny Restle. She shows how from a very early stage musical instruments were included in systematic classifications that were significant for the further development of the instrument as well as for the concrete practices of virtuoso control, and whose influence on notions of the instrument deserve consideration beyond the music sphere. Therefore the explanation of the systems of Praetorius, Mersenne, and Kircher are combined with terminological and historical reflections. For example, she considers it noteworthy “that Praetorius takes the term ‘instrument(s)’ for granted. He uses it without
Introduction
XVII
reflecting on whether other ‘tools’ (e.g. medical or technical) could possibly be understood by this term” (260). In the case of Athanasius Kircher however, we find diverse links to anatomy (the human voice) as well as to the acoustic experiments of physics. Andreas Meyer provides an original study of the key role played by musical instruments in a new understanding of scientific and artistic instruments by showing how African instruments were characterized in European sources. Particularly noteworthy here is the self-assured air, which seems to be motivated, not least, by the increasing circulation of ingenious instruments. Therefore, the study deals not only with the exotic instruments described, but also with the “language and tone of the sources” (277). References to the theater, to the spectacular aspects of experimental presentation, to factors of illusion and mechanical phantasmagoria are gathered and explained by Jan Lazardzig from the machine books of the time, starting with the notion “that the spectacle . . . is essential to the idea of the machine in the seventeenth century. Relating to the audience is necessary in bridging the gap between illusion and utility, and allows the machine to become an object of admiration and therefore be guaranteed to ‘function’” (153). Against this background, observations on the relationship between functionality and admiration can be situated. Theatrical machines offered manifold possibilities for the practical production of instrumental knowledge. In this contribution, a political dimension is also noteworthy, connected with the term ‘state machine’ as well as the related dimensions of audience and public. In Florian Nelle’s study on the instrumental revelation of new worlds, the theater with all its dramaturgical, technical, and scenic means is also placed at the center. If the conscious functionalizing of the theatrical world of appearance for the production of astonishing effects can be described, in part, as an instrumental strategy, then scientific instruments also occasionally prove to be the source of enigmatic signs and wonders. In this essay, theater and instrument are cross-referenced in order to trace the production of the wonderful as a side effect, typical for the time, of knowledge production. Here not only instrumental alienation effects are considered. One might even speak of a “poetics of the instrument” (66). Starting with a comparison between the theatricality of the ideal museum of Athanasius Kircher and the exercitia spiritualia of St. Ignatius of Loyola, Angela Mayer-Deutsch analyzes the effective practices of the instrumentalization of pictures against the background of Jesuit experimental arts, contemplation, and rhetorical persuasion. It becomes
XVIII
Helmar Schramm
clear how the systematic development of a religious pictorial rhetoric is bound up with the shaping, specification, and more precise definition sharpening of the term ‘instrument’ in the field of tension between science and art. If on the one hand religious techniques of persuasion are bound up with the experimental arts (where above all problems of optics and the magnetic powers of attraction and repulsion dominate), then on the other hand it is also a question of the theatrical dimensions of an art that is meant to delight, surprise, and astonish the observer. Therefore, it can be imagined that the stimulus for the instrumentalization of images in the Musaeum Kircherianum was probably drawn from the “mystery plays and the propagandistic Jesuit theater” (250). If one wants to examine the culture of the instrument in the seventeenth century focussing explicitly on correlations between art and science, then the alchemistic laboratory whose implements, tools, practices and materials reveal centuries-old traditions provides an interesting point of reference. Not merely because “the ‘Entstehungsherd’ (Friedrich Nietzsche) of experimental practice whose consequences reach into our own times” (201), but also because alchemistic experimenting was permeated with poetic elements, gradually rejected in the course of the century to make room for the founding of modern chemistry. Therefore, it is a highly dynamic, conflict-laden field of investigation that is assessed in Gerald Hartung’s contribution. At the center are the alchemistic activities of Johann Joachim Becher, arising from the field of tension between project creation and experimental art. Particularly relevant to the question concerning the practical and terminological profiling of a historical understanding of the instrument is a description of the laboratory that is not exhausted by a presentation of all possible measuring implements, apparatuses, combustion furnaces, and test materials, but which also includes an alphabetum minerale, implicating a natural-philosophical doctrine of principles and an ordering schema of moral elements. Nicola Suthor productively discusses the relationship between hand and instrument from an art-theoretical perspective, through an investigation of the brush as the decisive instrument of painting, as the “mechanical link between hand, paint, and canvas” (106). For this, she is supported on the one hand by exemplary self-portraits from the seventeenth century showing the artist at work, and on the other by attempts at a theorization of the brush as pars pro toto of art. In a very complex presentation she manages to interrelate questions of painting technique, virtuoso control of the instrument, and aesthetic concepts. It gradually becomes clear that the suppression of artisanal factors of pictorial pro-
Introduction
XIX
duction corresponds to the emergence of a definition of art characterized by the detachment of ars from craft, and the understanding of painting as part of the artes liberales. By thoroughly examining the relationship between spirit and body, scholar and artisan, interesting connections with other contributions in this volume frequently emerge. Frank Fehrenbach also deals with the relationship between hand and instrument. His contribution, against the background of an informed analysis of Leonardo’s drawings, highlights “one of the smallest but nevertheless most sophisticated tools of the human dominion over nature – the draftsman’s instrument” (81). Leonardo’s technical drawings, relating to questions of hydraulics and military technology as well as problems of mechanics and aeronautics, are bound up with the development of instrumental culture in the seventeenth century not merely as a result of their historical precedence. Drawing emerges as a cognitive medium preceding all invention of machines and instruments. Only in its active relation to drawing pen and drawing does the hand become an ‘instrumentum instrumentorum.’ Frank Fehrenbach adroitly describes related developments in the fifteenth and sixteenth centuries, without knowledge of which many aspects of the instrumental culture of the seventeenth century would remain obscure. Strongly divergent evaluations of image-formation by means of optical devices cannot really be understood without reference to a centuries-old tradition of spiritualization, and the production of dramatic effects that suggest the appearance of divine powers. Barbara Maria Stafford follows such traces, and reveals in her extremely vivid and perceptive study important contexts that explain Protestant uneasiness concerning visual representations in post-Reformation Europe. A methodological peculiarity of her approach consists in the attempt “to demonstrate the elaborate intersections of theater, theology, technology, and aesthetics” (128) by means of which transformations of visual culture can be explained. The whole contradictoriness of the object of investigation emerges clearly, in large part since the use of projection technologies is characterized as a highly contentious battleground of religious, political, scientific, and artistic interests. This contribution is not limited to opening up a broad-reaching field of study, however. Instead the presentation culminates in a whole series of fundamental theoretical inferences on the instrumental conditions and medial origins of art. Don Ihde also considers projection technology, in the form of the camera obscura, though he does not focus on the optical devices and experiments themselves, turning his attention instead towards the “philosophic and epistemological direction” (387) of such instruments.
XX
Helmar Schramm
Considering the long early history of the camera obscura, he notes that art, with its aesthetic concepts, experimental practices, and imagery, was ahead of the new establishment of scientific work. In the seventeenth century, for Descartes, Locke, and others, the camera obscura became “the very model for the production of knowledge” (388). The thesis of the epistemological relevance of technologies as a model for the production of knowledge can also be seen in the continuation of developments up to the present. Martin Burckhardt considers a problem that must be of the greatest importance for all production of knowledge, namely appearance [Schein], i.e. the new quality of the production, administration, and instrumentalization of appearance in the seventeenth century. The original as well as productive nature of his approach is due to a double interpretation of the word ‘appearance,’ firstly as illusion, phantasma, projection, and secondly as bank note [Geldschein]. Central to his considerations is the establishment of the first central bank, the Bank of England, at the end of the century, which he understands as a necessary functional element of a “machine of central perspective,” where “a centralization of power in the actor playing the king, a centralization of force in the standing, uniformed army, a centralization of money in the bank note” (452) structurally overlap and complete each other. The attempt to grasp the situation of the seventeenth century is bound up with a survey of the European history of money. As a result of such a historical contextualization, the author concludes that the special achievement of the seventeenth century was “not so much in the creation of new instruments as in the systemic completion and implementation of the resulting figures of thought” (449f.). Precisely in this light, the setting up of the central bank and the accompanying instrumentalization of appearance [Schein] is shown as a significant case of great symptomatic value. Dieter Mersch’s contribution is also concerned, though from a completely different perspective, with the production of appearance, with highly charged interrelations in the construction of rationality and irrationality in early modern modes of presentation. His analysis concentrates on the geometric thinking so important in the seventeenth century, and particularly on an aspect that played “an important role in the development of art and visual representation from the sixteenth to the eighteenth centuries: the dialectic of central perspective and anamorphosis” (21). However, this study differs from those presentations on the history of anamorphosis which appear merely as a footnote to the art history of the sixteenth and seventeenth century. Here, a context is brought into play that extends from the beginnings of Euclidean geome-
Introduction
XXI
try to the broad-ranging establishment of non-Euclidean geometry in the nineteenth century. Within this broad framework, the extent to which “the dynamic of a culture cannot be written in a one-sided way, either from the viewpoint of technology, or from the viewpoint of the progress of mathematics; instead it shows that art and science play off each other and that they are interlocked in a complex way” (35) is impressively demonstrated. With her investigations into the convergences between science and art in the early modern period, Sybille Krämer turns expressly against the widely circulated conception of an instrument(ality) founded on a clear distinction between the technical and the symbolic, production and representation, construction and interpretation, the instrument and the medium. By contrast she highlights the significance of symbol-technical hybrids, which results in the thesis “that lasting shifts or even ‘leaps’ of a cultural dynamic touch on a – in cultural-technical terms – definable ‘instrumentality’ of just such symbol-technical mixed forms” (458). The concrete anchor of this analysis is the practical use of the mathematical zero, resulting in the emergence of interesting relations to the broadranging career of the central-perspective vanishing point. The consequences for the dimension of human modes of handling, forms of perception, and linguistic practices that result from the constellations of linear perspective and formal language are also worthy of note. If one considers the complex introduction and development of highly diverse fields of research in the seventeenth century, then the pioneering function of zero “as a ‘middle’ and ‘mediator’ which makes heterogeneous worlds relatable, and mutually translatable” (471) is emphatically underlined. The image-worlds of Robert Fludd published around the middle of the seventeenth century create for Olaf Breidbach an important point of reference for the connection between questions concerning the historical development of knowledge and media-historical considerations. Ideas of the logistics of research, notions of the instrument and instrumentalization are here confronted with a concept of the model, in whose light concrete aspects of knowledge production emerge. The model outlines an idea “that is not completely translatable into a language of science, into a formula, or into a complex sentence, and that cannot be explained by means of these formulations. As such, a medium for this idea is found in the model. The idea remains fixed to the forms of this medium, to what is technically possible and therefore also to the reality of this medium” (39f.). It is therefore a question of investigating the material conditions of knowledge production as well as the concrete localizations of experimental works. For such questions of localization, a growing aware-
XXII
Helmar Schramm
ness of problems is developed that is not only oriented towards the results of geographical or astronomical cartography, but also – and this is demonstrated using Robert Fludd’s image-worlds – towards an “architecture of the human,” (53) a “topography of the human” (56). Only with the inclusion of this dimension, connected with the consideration of technical-instrumental aspects of knowledge production, do media-historical observations acquire their heuristic value for the discourses of the present. Ludger Schwarte, in his contribution, first distances himself from positions that limit the notion of the instrument to mostly tool-like functions, discussing interactions between instruments and handling complexes as well as the instrumental formation of language and perception. However, the work on a definition of the instrument is only a starting point, providing the basis from which to advance to much more farreaching considerations. Using thorough research into the founding, circulation, and practical functioning of the theatrum anatomicum in the early modern period, a constellation of body, model, instrument is analyzed, which, from a historical as well as theoretical point of view, proves to be very inspiring and constructive. The paper shows how, since the middle of the sixteenth century, activities in the theatrum anatomicum were directed towards the construction of a universal human body, and how this body-model functioned concretely as an instrument of cognition. The next step is to question the consequences of the relationship between brain anatomy and visualization. The key thesis suggests that visualizations “affect our understanding of thought and the practices of the brain” (183). If Adorno and Horkheimer once spoke of an instrumentalization of reason, then, from the foundations set out here, highly interesting possibilities for a critical resolution (Aufhebung) and further development of this conception arise. Such anticipatory considerations are promisingly hinted at in the concluding section on the instrumental practices of the brain. The integration of instruments into concrete local conditions as well as the related “process of defining the sciences by the praxis of their representation” (224) is investigated by Gerhard Wiesenfeldt with the help of thorough archive studies of the University of Leiden in the period around 1700. The results that come to light here are very striking. Apparently obvious speculations on the order of knowledge, as revealed in the concrete givens of library, physics theater, anatomy theater, observatory, and chemical laboratory, are often, with closer observation, completely turned on their head: “Since at no point did what was actually achieved in terms of experimental work in the sites of representation, and which could have led to a thematic contouring of the disci-
Introduction
XXIII
plines, have an essential significance for the self-understanding of the disciplines represented” (233). This conclusion, also readable as a plea against hasty generalizations, shows to what extent the practical visible aspects of the experimental use of instruments are linked to a partly invisible architectonics of cultural boundaries that cannot be characterized merely by architectural and local givens, but additionally by political, social, economic, and disciplinary demarcations. It becomes clear that the architectonics of cultural boundaries are not to be understood as a rigid system, but as a dynamic process with its own rhythms, which frequently cross-fade with surfaces of representation. The in-depth contribution by Thomas F. Gieryn on the instrumentality of places of science and art can be understood as a link to the first volume of our series ‘Theatrum Scientarium,’ which is explicitly about scenes of knowledge in the seventeenth century. In this paper, the local economic and communicative conditions of two significant American schools of science and art are presented. The first deals with the New York School of Abstract Expressionists, while the second presents the Chicago School of Urban Sociologists in the course of its development. Admittedly, while reading it is difficult to get rid of the impression that this study is only partly anchored in the program of the current volume. By the end, however, many details which at first glance do not seem particularly relevant, lead to a far-reaching conclusion, which can be useful for the historical and systematic understanding of instruments, as well as for grasping the relation between instrument and hand(ling). The author encapsulates his case studies in the following summary: “Instruments used for scientific and artistic work have a place. They exist at some particular spot in the universe – bounded, spatially delimited. They are located amid assemblages of buildings, streets, tools, supplies, cafés, and maybe trees. Such assemblages, such places, themselves become instruments for the production of science and art” (394). Fundamental connections between hand and instrument, artisanal skill and instrumental armament are examined by Stefan Ditzen in relation to early microscopy. This study shows in a striking way how the imageworlds of microscopy are influenced not only by religious motifs, but above all by the completely concrete material conditions of knowledge production. The quality of the depictions not only depended on the cultivated instrumental gaze of the experimenter, but also on the given artisanal skills of the lens cutter or engraver. The economic implications of experimenting are also dealt with, high-grade lenses being in no way easy to acquire, so that in most cases one fell back on completely deficient implements. Such difficulties alone can explain how with the
XXIV
Helmar Schramm
circulation of certain illustrations in publications the object of investigation actually seen by the experimenter frequently becomes over-layered with already existing pictorial models. In particular, the richly illustrated Micrographia by Robert Hooke was repeatedly used by experimenters to prove in a deceitful way the success of their own microscopy. Such traces of plagiarism are shown very beautifully in this study. Using an illustration of a mite from Hooke’s famous work, we discover “how constant such a model of the seventeenth century could be” (346). Such conflicts between the microscope as an instrument and the particular modes of instrumentalization of microscopic images show to what extent the production of knowledge in one way or another can be related to the emergence of epistemological obstacles and the limits of the gaze. An excellent supplement to this study of the beginnings of the microscope is provided by Jochen Hennig’s study of the instrumental conditions in the images of modern scanning tunneling microscopy. However much procedures of image-production directly correspond with the instrument of production used, and no matter why in particular phases of the work the implement used loses its status of instrument and necessarily becomes an object of investigation itself, this dynamic of research as practical production appears cogent, and encourages the investigation of similar questions within the history of science as well. What is established here concerning the production and publication of scanning tunneling microscopic images could also be applied to countless historical instruments, test setups, and work processes: “Many scientists also developed particular skills. These did not become the subject of systematic publication” (354). Bruno Bachimont also ventures to create a vigorous link between the legacies of the seventeenth century and current developments. His contribution culminates in the Turing machine, indeed as a result of a representation which deals with aspects of a history of tools of thought in the narrow sense of the word, that is, of cultural-historical shifts of accent concerning the instrumentation of signs, which in their material concreteness are integrated into extremely varied contexts, and thus occasionally provoke instructive disturbances. To the extent that computer science is understood as a practical automatizing stage in the evaluation and inference of formal systems, it is also valid to give “the circumstances of computational tool use” (365) due consideration. This contribution achieves its strength through the skilful linking of historically concrete, localizable cases with long-term developments whose transformative power can often only be perceived with historical distance. In this light,
Introduction
XXV
it is also legitimate to let important factors of the instrumentalization of thought in the seventeenth century culminate in reflections on the Turing machine and computational intelligence, without running the risk of lapsing into a teleological approach to history writing. If the history of the arts and experimental sciences in the seventeenth century holds completely new impulses, then since the beginning of the nineteenth century at the latest their continued developments achieved a dynamic that led to a radical history of division between art and science as well as the humanities; this in turn had a considerable influence on the new understanding of the terms ‘experiment’ and ‘instrument.’ In this light, H. Otto Sibum’s contribution is of critical importance in the framework of our publication. Starting with Moritz Hermann Jacobi’s construction of a spectacular machine in the nineteenth century, of a ‘physical perpetuum mobile,’ an electromechanical machine, which was also understood by its inventor as a “motor of civilization” (284ff.), this study reveals deep shifts in the value-structure of the sciences, the status of the experimental scientist, of the learned writers as well as the whole logistical and experimental structure of experimentation. Here again the focus is necessarily turned towards the relationship between hand and instrument, which in the course of the eighteenth century, due to a prevailing dominance of written culture, could not hold a prominent position. Now however concrete processes of handling and production, against the background of industrial machinery, can no longer be ignored as factors that shape knowledge. As the experimenter and engineer gain in influence and esteem, the awareness of problems of the culture-forming constellation of hand and word also acquires new emphasis. Also affected by this is the situating of instruments in the field of tension between art and science, where increasingly the manual skill of the experimenter and the masterly control of the instrument by the musician meet in a reevaluated notion of virtuosity. Hans-Jörg Rheinberger’s contribution proposes fundamentally to address and work through the question of the relationship between instrument and experiment. Using the example of a whole series of instruments that played a role in the biosciences of the nineteenth and twentieth centuries – ranging from the microscope through physiological apparatuses, model organisms and X-ray crystallography up to radioactive isotopes and scintillation machines – he shows “that the use of instruments in the biological sciences brings with it the necessity to create very differently shaped, generally opaque intersections between the possible objects of inquiry and the instruments that become involved in the investigation” (2). The interface as surface limit between apparatus
XXVI
Helmar Schramm
and object becomes the real pivot of a dynamic conception of the instrumental, from which inspiration for the new accentuation of historical research undoubtedly also results. By marking the interface as a decisive epistemological problem zone, Hans-Jörg Rheinberger draws our attention to the relevant limits of the micro-arena that also result from the specific character of epistemic objects and instruments: ‘small limits,’ without the observation of which the architectonics of cultural boundaries would remain a great riddle. In very practical terms, the interface can also be understood as a concrete place “where the hand of the instrument maker shades and grades into the hand of the experimenter, and even in the age of the industrial production of research technologies they remain the loci of handicraft” (ibid). Peter Galison and Lorraine Daston consider the fundamental conditions for the production of knowledge by focussing on the problem of the ‘many-headed knowers,’ i.e. the production of knowledge in communities and networks. Here it is not merely a question of the standardization of locally and technically differing instruments. “The observers themselves must be socialized and standardized in order for the networks to cohere and endure. This is the juncture at which ethos met epistemology” (300). With the appearance of ‘many-headed knowers,’ not only the quantity of gathered data, but also the whole style of thought changes. As the relevance of this problem has been raised consistently since the time of Descartes, it is only logical that this study analyzes significant cases from the late nineteenth century. Using the Carte du Ciel, the Internationale Gradmessung, and the Transits of Venus, questions of the production and coordination of operational instruments, questions of searching, discussing, and inventorying sources of error, the physical and psychological synchronization of observers, and the logistics and economy of working are analyzed. The way in which the combination of the achievements “of local observers into a kind of Leviathan or super-observer” (303) profoundly corresponds to the political and social contexts of research is also convincingly presented. In addition, the relation between instrument and hand(ling) emerges in a particularly productive way. Finally Georges Didi-Huberman shrewdly considers the systematic re-evaluation in the early modern period of a new science supported by the instrument and the corresponding ideal of the systematic dissection, inventorying, and controlling of time and space indirectly, by calling on Henri Bergson’s philosophy as a kind of precision instrument of critical inventorying. Using this procedure, not only are facets, traces, and images of the situating of the instrument in art and science around 1900
Introduction
XXVII
examined, but flashbacks to the Cartesian period are also suggested – was Bergson’s experimental philosophy nothing other than a radical counterpoint, understanding itself as rejection of any blind trust in the instrument of concise concepts and hierarchical systems? In this light, his critique of “cinematographic illusion” is in no way merely directed at the discussions of scientific and artistic strategies of his time, but aimed beyond this at the decisive foundations of the construction of theories and the production of knowledge as established in early modern Europe. “The ‘cinematographic illusion’ . . . consists of believing one sees everything and of manipulating this everything seen like an extension which is divisible at will, a quantity integrally geometrizable, an easily instrumentalizable vital energy” (435). In his search for flexible, supple concepts, and in addition, in his attempt to replace the terminological instrument with images – images understood as “the molding of things (as the plaster works of Rodin, in Bergson’s time, retained the singularity of the least element of his formal vocabulary)” (425) a fundamental awareness of crisis is articulated at the threshold of the dawn of a completely new age of physics, science, art, and their related instrumental foundations.
HANS-JÖRG RHEINBERGER
Intersections: Some Thoughts on Instruments and Objects in the Experimental Context of the Life Sciences* Since the seventeenth century, scientific instruments have been seen as the emblems and signposts of experimental science. Over and over again, they have been described and hailed as ideally transparent media that either help to prolong, reinforce, and super-elevate the senses or to isolate, purify, and quantify experimental perceptions. Historical literature on instruments has been proliferating over the past decades; it shall not be reviewed here in detail. Importantly, historians of science have thoroughly challenged and questioned the assumption of instrumental transparency. In particular, many case studies have shown that as a rule, instruments do not work by themselves and do not generate evidence by themselves. Rather, they are embedded in historical and local contexts of skill and application, without which their production and their efficacy cannot be understood; outside of which their functioning cannot be granted; and which also determine the range of their circulation. 1. Epistemological Preliminaries I would like to use this essay in order to raise the basic question of the relation between instrument and experimental object. What configurations can it assume? Where does the experiment take place? Is it at the instrument, with, before, within the instrument? How, accordingly, can the particular epistemic value of an instrument be determined? In what follows, I will not present another case study and not reconstruct another local historical context. Rather, I would like to discuss a confined but fun*
I would like to thank Sven Dierig, Peter Geimer, Henning Schmidgen, and Laura Otis for a discussion of the text and for helpful and valuable suggestions. An earlier version of this text appeared in the Report 2002/03, Max-Planck-Institute for the History of Science, Berlin 2004.
2
Hans-Jörg Rheinberger
damental epistemological problem. I would like to show that the use of instruments in the biological sciences brings with it the necessity to create very differently shaped, generally opaque intersections between the possible objects of inquiry and the instruments that become involved in the investigation. Such intersections mark the contact surface between the apparatus and the object. In examining a series of instruments that came to be employed in the life sciences of the nineteenth and twentieth centuries in particular, I will show how these points and planes of intersection between the living and the non-living, between the organism and the technical apparatus were configured. The investigative value of an instrument depends on the shape of such intersections; they decide whether a particular instrument and a particular object can be brought together at all and bound into a fruitful analytical constellation. Consequently, such constellations have been the locus of particular artisanship and attention. The intersections are the places where the hand of the instrument maker shades and grades into the hand of the experimenter, and even in the age of the industrial production of research technologies, they remain the loci of handicraft. In the context of the development of new research technologies, the exploration of intersections between object and instrument is often at least as important as the technical implementation of the scientific principle embodied by the instrument, although neither need necessarily have an intrinsic, theoretically motivated relation with the other. Work on the intersections may even at times develop into a separate industry, as was the case, for instance, with electron microscopic specimen preparation. Unfortunately, the work at the intersections between the instrument and the object of investigation has not always received the historiographical attention it deserves. It is this particular, rather narrowly circumscribed question that I want to address here. It stands, however, in the context of a wider epistemological problematic, which I therefore want briefly to explicate first. This wider question concerns the relation between epistemic objects and the technical conditions of their manipulation in the framework of experimental systems. In describing experimental systems I have, on several occasions, pointed to the fact that the productivity of such systems essentially depends on a well-balanced dialectic between epistemic and technical things. The technical things bound and confine experimental systems. They constitute a more or less rigid framework of conditions, and at the same time, they determine a scope of action in which an epistemic object can unfold. My general claim is that instruments receive their epistemic meaning only in relation with and as a part of experimental systems. Taken by themselves, they are epistemically inde-
Intersections
3
terminate. Although they can be viewed as embodiments of certain theories or concepts, that is to say, with Gaston Bachelard, as “reified theorems,” they are not knowledge generating instruments in themselves.1 As a rule, instruments enter as technical things into an experimental arrangement, as the identity conditions of an experiment, but they can also become epistemic things, if in the course of their use they generate unexpected questions. In the development of research technologies we often observe that the crafting of an instrument goes hand in hand, and is inextricably linked, with the process in which an epistemic object takes shape. The intersections between the instrument and the object thus form a particular problem zone of the articulation between epistemic and technical things. As far as our narrower question is concerned, we should consider the fact that in biological experimentation, the intersection between the object of investigation and the technique of its representation or measurement is always also, very concretely, a boundary between an organic body and an inorganic entity. It is a place where life and technique confront each other, and since, as a rule, the living part is wet and the technical part is dry, their encounter requires particular precautions. The success of a biological experiment depends on mastering this transition, that is, on shaping the joints intended to make the wet and the dry, the soft and the hard, the fluid and the solid compatible. In the long run, the productivity and the sustainability of biological experimental systems is determined by the way this boundary is dealt with. Such boundaries, such intersections will be characterized in this essay. I would like to demonstrate their multiform shapes by describing a few examples more precisely, and while so doing touch upon a few epistemological questions of a more basic nature, all of which are concerned with instruments and experimentation. They are all concerned with the particular materiality that characterizes epistemic objects in the life sciences. Therefore it is not the architecture of the great boundaries, such as the articulation between the sciences and the arts, the demarcation lines between disciplines, or the relationship between the sciences and other cultural formations that is in the foreground of attention. What we are concerned with is rather the small boundaries: those soft lines of demarcation and partition between what is to be taken as biological nature and what is to be rejected as artifact. As soon as scientists embarked on the endeavor of turning the internal constitution of organisms inside out, they were confronted with the problem of providing the rele1
Gaston Bachelard. Le rationalisme appliqué. Paris: P.U.F., 1998. 103.
Hans-Jörg Rheinberger
4
vant precautions for their boundary work. The life sciences have been haunted and persecuted by the question of the legitimacy of reconfiguring these boundaries since the time they took up the challenge of exploring organic nature in its inner structure, that is, of putting hands on the material conditions of its functioning. We will see that the drama of drawing these small boundaries not only pervades all the life sciences, but is also decisive for an understanding of the major divides between disciplines and scientific cultures. 2. The Microscope2 I will start with microscopy. With the emergence and diffusion of the microscope since the second part of the seventeenth century, the question of the object of observation poses itself in a new manner in natural history. A new form of specimens of small dimensions emerges that corresponds to the new technology of magnification. Microscopy is a good example of the general need to correlate forms of observation linked with new instruments with the state into which things have to be brought in order to be visualized by these means. On the one hand, the things prepared for the lens cannot themselves be seen in the process of preparation, at the very least not in the detail on which the future success of the observation depends. Their preparation escapes the capturing eye; it is only the gaze through the microscope after the fact that will decide whether the preparation was successful. This forces the preparators to direct their attention to the regulation of the process of preparation. Since the object of inquiry remains inaccessible to direct observation of the eye, the procedure must, in one way or another, function blindly. It is therefore not by chance that the scientific literature pays ever more attention to the techniques of preparation and describes them in great circumstantial detail. It is in the nature of fresh specimens that they must be newly supplied for each observation. But how can one ensure the orderly repetition of the procedure? In view of this question it does not come as a surprise if, to take just one example, Matthias Jacob Schleiden, in his debate on the fertilization of plants with Franz Ferdinand Meyen, describes his proceedings in painstaking detail. In the second part of his Botany as Inductive Science (1846), we read: 2
This section is based on Hans-Jörg Rheinberger. “Präparate – ‘Bilder’ ihrer selbst. Eine bildtheoretische Glosse.” Oberflächen der Theorie (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1,2). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2003. 9-19.
Intersections
5
Here I would like to add a few words on the manufacture [Darstellung, he says] of such preparations. If the buds of the seeds do not lie very tightly enclosed and immobile in the ovary, I prepare them free, then take them between forefinger and thumb in such a way that I can sever them exactly in two halves with a sharp razor blade . . . The two halves thus acquired I lay one after another, with their cut surfaces pointing toward the thumb, again between the two named fingers, and then I use a razor blade to cut off as fine a section as possible from the sectional plane. – Then I place these two disks under the simple microscope and, with the help of fine needles and small knives, lay bare the relevant parts, if they are not, this case always being the best, already exposed by the cut itself.3
In this “exposure by the cut itself,” in the exertion of a minimum of additional manipulation, the freshly prepared, wet botanical specimen finds its master. But attempts to render it durable soon follow, for, in the last instance, only dry preparation ensures the permanence of the viewed object and with it, the possibility of retrievability and comparison with other preparations.4 In that way, differential reproduction of specimens becomes possible. But since making them durable requires additional intervention, the question of what is nature and what artifact in view of the preparation acquires a special epistemological urgency. For here, in contrast to macroscopical observation, visual comparison with the ‘living’ counterpart is impossible. Microscopic preparations are therefore epistemically highly laden objects of knowledge. It is only logical that the methodical critique of the knowledge practices of the life sciences in the latter part of the nineteenth century has crystallized to a considerable degree around these preparations. In addition, it is a defining feature of microscopic preparations that they reduce the objects fixed in them to two dimensions. They flatten them out. This is a necessity grounded in the functioning of the apparatus whose focus is precisely on a plane. The microscope, with its imaging capacity, not only remains on the surface, but remains bound to a flat plane, and the flattened object is realized by means of the “cut.” In close connection with the establishment of cell theory, the tissue section becomes an emblem of animal and plant microscopic morphology in the nineteenth century. The zoologists, botanists, anatomists, physiologists, and microbiologists of the second half of the nineteenth century revolutionized the craft of the microscopic cut by mobilizing the newly ac3
4
Matthias Jacob Schleiden. Die Botanik als inductive Wissenschaft behandelt. Zweite, gänzlich umgearbeitete Auflage der Grundzüge der Wissenschaftlichen Botanik. Zweiter Theil: Morphologie, Organologie. Leipzig: Engelmann, 1846. 370f. Cf. Jutta Schickore. “Fixierung mikroskopischer Beobachtungen. Zeichnung, Dauerpräparat, Mikrofotografie.” Ordnungen der Sichtbarkeit. Fotografie in Wissenschaft, Kunst und Technologie. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 285-310.
6
Hans-Jörg Rheinberger
quired powers of inorganic and organic chemistry, of acids, and of dyes. Various fixing, dying, and hardening procedures rendered the preparations lasting, made new contours visible, soft things cuttable. These procedures not only mark the point of severance between the organic body and the optical apparatus, they also form and shape it. This line of fracture is the precarious point where the epistemic object and the instrument become intertwined. Around it, the experimental systems of microscopy accrete. It is here that the organic and the technical engage in mutual interaction. Here, at the very point of magnification, the decision is made about the way in which the specimen will enter into the picture. Without the procedure employed in trimming specimens, it would not have been possible to exhaust the potentials of the new lens technologies developed in the course of the nineteenth century. With them, the locus of magnification was shifted from the wet side to the dry side. The fresh cut of the razor blade was subjected to a changing bath of chemical reactions. The microscopic things themselves, treated with acids, dyes, and fixers, embedded and welded between object carrier and covering glass, started to fill the cases and boxes of microscopic archives with a new form of epistemic objects rendered durable. What was especially difficult to cut, what could not be taken between thumb and forefinger, was embedded in resins. Microtomes were developed that were able to cut slices of a hitherto unknown thinness. Using this whole arsenal of new techniques, the microscopists transformed the work of preparation into a space within whose coordinates they came to play out a continued dialectics of fact and artifact – to take up an expression of Bachelard again in this context. The more clearly and sharply they tried to make something visible, the more they brought it near to that boundary at which one can no longer decide what one has conserved: the object or the means of its objectification. In the borderline case, and as Peter Geimer has argued using the example of scientific photography, the preparation comes to represent the preparation technology itself.5 All representation in research revolves around such cusps. Instead of problematizing them as cognitive traps, we should see and understand them in their positivity: as driving forces, as engines that maintain an epistemic dynamics which brings the object of interest into a form that can, if necessary, be transcended. It is in the nature of epistemic objects in general that they can become outstripped. They are 5
Cf. Peter Geimer. “Was ist kein Bild? Zur ‘Störung der Verweisung.’” Ordnungen der Sichtbarkeit. Fotografie in Wissenschaft, Technologie und Kunst. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 313-41.
Intersections
7
obtained in a recursive fashion, and they remain relevant for research just as long as the work of deconstruction can go on with and around them. It is therefore not by chance that microscopy occupies a decisive place in a developing epistemology of error, as Jutta Schickore has shown in her investigations into the discourse of error in the microscopic life sciences of the nineteenth century.6 Around the work of microscopy, the methodological consciousness of a science crystallizes that constantly moves along the boundary between the visible and the invisible – which is at the same time a boundary between the living and the dead – and that, in order to overcome this boundary, has to subject its potential objects to ever new interventions. 3. Physiological Apparatuses The apparatus-based physiology of the nineteenth century exhibits another picture of the intersection between the organic body and the technical gadget. There are plenty of investigations into the experimentalization of nineteenth-century physiology that cannot be considered here in detail. The only thing that is of interest in the present context is the configuration of that point at which the organism – or parts of it – and the instrument come in contact with each other. We can take the kymographion of Carl Ludwig as an example (fig. 1). In the context of this registration device the point of intersection takes on the form of a genuine lesion, if not mutilation. The apparatus developed by the Leipzig physiologist Ludwig allowed the measurement of blood pressure in a living animal. In the process, a “communicator” connected the open wound of the animal with the curve registration part of the machine. In his work on nineteenth-century laboratory physiology, Sven Dierig has argued extensively that the success of the instrument, that is, of obtaining reliable blood pressure curves, depended critically on the form and the properties of this interstitial piece. All possible forms and material means of connection were tried out in collaboration between physiologists and instrument makers, until finally the mercury pressure gauge succeeded. Here, the liquid metal took up the pressure of the arterial blood through a glass cannula inserted into the artery; at the other end of the U-bend, the rises and falls of the meniscus of the metal were transmitted to a pencil that transformed them into traces on paper mounted on a rotating 6
Cf. Jutta Schickore. “(Ab)using the Past for Present Purposes: Exposing Contextual and Trans-Contextual Features of Error.” Perspectives on Science 10 (2002): 433-56.
Hans-Jörg Rheinberger
8
Fig. 1: Blood pressure experiment with a Kymographion.
cylinder.7 The medium that transformed the organic movement into a technical movement had to be in resonance with the manifestation of life that was being investigated. In this case, the conductive medium was a fluid that reacted to pressure. In experiments with nerves and muscles, however, electrical circuits formed the mediating connections. In EtienneJules Marey’s apparatuses for measuring gait, investigated in detail by Andreas Mayer, it was a compressible rubber bladder that formed the contact surface between the foot and the soil.8 In such experiments of a technologically highly equipped physiology characteristic of the second 7
8
Cf. Sven Dierig. Wissenschaft in der Maschinenstadt. Emil Du Bois-Reymond und seine Laboratorien in Berlin. Göttingen: Wallenstein, 2006. Cf. these and similar instruments online at: http://vlp.mpiwg-berlin.mpg.de (address of the Virtual Laboratory for Physiology). Andreas Mayer. “Autographien des Ganges. Repräsentation und Redressement bewegter Körper im neunzehnten Jahrhundert.” Kunstmaschinen. Spielräume des Sehens zwischen Wissenschaft und Ästhetik. Ed. idem and Alexandre Métraux. Frankfurt a.M.: Fischer, 2005. 101-138.
Intersections
9
half of the nineteenth century, the organism becomes an element in a technical construction in which everything depends on the seamlessness of the joints. Such frameworks not only serve to measure certain manifestations of the life of an organism; in borderline cases they also sustain these manifestations. Peter Geimer has pointed to the paradox that with these experimental hybrids, these cyborgs of the nineteenth century, the question of life or non-life, the question of where nature ceases and technology begins, is no longer answerable in an unequivocal manner.9 Because, at least in extreme cases, the organism whose life manifestations are analyzed and measured is no longer able to live outside the apparatus. It lives exactly as long as the machinery turns. During the nineteenth century, the major organic circuits such as respiration, blood circulation, and nervous conduction were transformed into objects of analysis in such a way that the organism was transformed into an organic element, even at times an organic switch, in a technically determined circuit. For each particular organic function, the junction was made by a corresponding mechanical equivalent of that function. 4. Model Organisms At this point, I would like to move to the other extreme and talk briefly about a different, counter-intuitive ‘instrument’ that has become characteristic of the life sciences of the twentieth century. It is the organism itself as a model. As Robert Kohler has forcefully argued for the pet of classical genetics, the fruit fly Drosophila melanogaster, model organisms function not only as exemplars, but also as ‘instruments’ of research.10 What does that mean epistemically, however, and is this manner of speaking more than a metaphor? In order to be able to answer this question, it is useful to come back once more to the concept of the experimental system. As an instrument, the model organism forms part of the technical conditions under which an epistemic object assumes its contours. Staying with the example of classical genetics, we can state that in the context of working out gene maps, Drosophila mutants do function less as epistemic objects than as tools with which genes – the epistemic entities under consideration – can be localized and their places fixed on chromosomes. And indeed, many of the Drosophila mutants 9 10
Cf. Peter Geimer, ed. Untot. Verhältnisse von Leben und Leblosigkeit. Preprint No. 250. Berlin: Max-Planck-Institut für Wissenschaftsgeschichte, 2002. Cf. Robert E. Kohler. Lords of the Fly. Chicago: University of Chicago Press, 1994.
10
Hans-Jörg Rheinberger
Fig. 2: Fruit fly mutant with ‘ski’ wings.
identified in Thomas Hunt Morgan’s laboratory were not interesting in themselves, but only as mapping markers. Taken by themselves, mutants of eye pigmentation and morphological mutants that were used early on as instruments in this sense are monsters (fig. 2), but as tools they are interesting in just this form: not because of the specificity of the defect, but because of the chromosomal location of the genes presumed to cause it. They took on the character of epistemic objects only decades later in the context of biochemical and developmental genetics, when the processes that underlie these features themselves became objects of investigation. Model organisms as instruments are peculiar in that they dispense with the problem of the intersection between organism and instrument. They are organisms turned into instruments. Exactly this makes them so powerful as tools. Here the instrument is made principally of the same organic material as the object of investigation.
Intersections
11
5. Test Tube Experiments and the Ultracentrifuge The question of intersection arises in a completely different manner with the biochemical experiments of the twentieth century. Biochemical test tube experimentation was instituted by Eduard Buchner’s efforts to obtain a cell-free alcoholic fermentation enzyme shortly before the turn of the century. What happens in the biological test tube experiment? Here we observe a new displacement of boundaries. It came to express the problem of what is biological nature and what is artifact in the process of investigation in the form of a special linguistic dichotomy. From now on, biologists distinguished between in vitro and in vivo experiments. As the expression betrays, with the in vitro experiment a vitreous envelope is created which replaces the walls of the cell, the wraps of the organism. What Claude Bernard termed the “milieu intérieur,” the inner environment of the organic processes, is replaced by a chemo-technical milieu that at the same time opens a new analytical space; a space turned inside out, tipped over, and tuned for new connections. The in vitro experiment, which for biologists was counter-indicated for a long time in view of the specificity of biological organization, develops its dynamics by allowing scientists to isolate particular organic reactions and their carriers and to represent them separately. One might be inclined to say that the spaces of intersection are now inserted between the parts of the organism itself, and become vitrified. The price that has to be paid is yet another radicalization of the question of what is being observed. Is it still a biological function, or has it shrunk to a chemical process? Is it something going on within an organ, or something created in the test tube? The question of how the results of an in vitro experiment can be reconfined in the space of the organism, of how they can be localized in the living, becomes the decisive question for a biochemistry that still aspires to understand itself as biological chemistry. The ultracentrifuge is an instrument that played a particular role in this context. Developed in the 1920s by Theodor Svedberg, it came to be used in the 1930s for the separation of cellular components and, as Angela Creager has described in detail, for qualitatively characterizing and quantitatively isolating viruses.11 The tube inserted in the rotor of the centrifuge and oriented in the gravitational field becomes a container in 11
Cf. Hans-Jörg Rheinberger. “Vom Mikrosom zum Ribosom: ‘Strategien’ der ‘Repräsentation’ 1935-1955.” Die Experimentalisierung des Lebens. Ed. idem and Michael Hagner. Berlin: Akademie Verlag, 1993. 162-87; Angela N.H. Creager. The Life of a Virus. Tobacco Mosaic Virus as an Experimental Model. Chicago: University of Chicago Press, 2002.
12
Hans-Jörg Rheinberger
which the contents of homogenized cells reorganize themselves according to a single parameter: the molecular weight of their constituents. Centrifugation allows investigators to decompose the cell sap in the rotor tube into sections separated from each other by more or less sharp boundaries. If the tissue is homogenized, that is, if the cells are broken up, the morphological and functional relevance of the centrifugal bands must subsequently be evaluated by reconnecting them to in vivo conditions through additional test systems involving whole cells. These checks and balances may be of a histological nature. A particularly elegant example of such a feedback is the centrifugation of intact cells. But the fractions can also be subjected to microscopic inspection under conditions similar to those described in section one. The tests may also be in vitro experiments, such as, for example, an enzyme assessment. In any case, however, the centrifugal partitions participate in the dilemma which haunts – and drives – all modern experimental biology, namely that of assuring itself of the boundary between what is still organic and what is no longer organic after that boundary has already been transgressed. These acts of transgression therefore always happen in the anticipation of a possible recursive assurance whose success no one can predict in advance. 6. X-Ray Crystallography X-ray structure analysis is another molecular technique that was first applied to polymers in the context of organic fiber research in the 1930s. Here the intersection with the biological object of investigation takes the form of a particular physical object, a crystal (fig. 3). What cannot be crystallized does not exist as an epistemic object for this technology. X-ray crystallography contributed decisively to a biophysical view of the basic structures of life. In the eighteenth century, the crystal analogy was already a favorite metaphor, and it found multiple uses in the nineteenth century. But only in the twentieth century did it materialize in the form of macromolecular biocrystals. During the twentieth century, the crystallization of biomolecules such as nucleic acids and proteins has decisively contributed to the breakthrough of a new view of biological order. On the one hand, X-ray crystallography led to the image of the iterative, double helical structure of DNA, in whose nucleotide sequence hereditary information is stored. On the other hand, it helped to visualize the three-dimensional structure of proteins as the translation products and functional correlates of nucleic acids. Soraya de Chadarevian has pointed to the fact that in order to translate back the mathematical Fourier world
Intersections
13
Fig. 3: Crystals of ribosomes (70S) from the bacteria Thermus thermophilus (a) and Bacillus stearothermophilus (b).
that is engendered at the intersection between the organic molecule turned crystal and the X-ray, the crystallographers had to create a parallel world of macroscopic models that helped to project the molecular, crystallized point of intersection back into the world of three dimensions.12 7. The Electron Microscope Twentieth-century electron microscopy has once more radicalized the process of microscopic specimen preparation already described. Nicolas Rasmussen – in the case of the United States – and Bruno Strasser – in the case of Switzerland – have characterized in detail the precautions and frictions under which a technology that had been developed in the 12
Cf. Soraya de Chadarevian. Designs for Life. Cambridge: Cambridge University Press, 2002.
14
Hans-Jörg Rheinberger
context of the material sciences, and that had not initially been intended for biological application, was made suitable for biological objects.13 On the one hand, electron microscopy has forced the preparation of ultra-thin sections. Penetration by the electron beam could only be guaranteed if the thickness of the specimens was dramatically reduced. New embedding procedures, microtomes with minimalized feed, and electron dense ‘dyes’ for contrast enhancement contributed to the formation of a new field of research undertakings that developed in parallel with the instrument itself, the electron cannon. The biological material itself had to be treated in a such a manner that it resisted the harsh conditions to which it was exposed in the microscope, at least as long as an electron shade was generated and recorded. The durability of an electron microscopic biological preparation is almost inevitably restricted. The specimen has to be exposed to a strong vacuum and is consumed by the bombardment of the electron beam. In contrast to the procedures of light microscopy, the interaction with the instrument during the screening process and at the moment of picture formation is so strong that the preparation itself tends to be destroyed. Strangely enough, here, at a new peak of specimen preparation technology, a point is reached where the object of investigation again – as was the case with the fresh cuts for light microscopy – becomes transitory. It is lost in the act of making it visible. As a result, the electron density and the electron resistance of the material brought under the beam become the decisive parameters for modulating the surface of intersection between the instrument and the epistemic object. This is also the point where contrast enhancement comes in. One ‘dyes,’ for example, with electron dense salts containing heavy metals. One of the most remarkable modulations of the intersection plane between the electron cannon and the biological material consists in converting the organic surface into a metallic replica (fig. 4). The specimen is covered with a coat of metal evaporated from a metal source at a certain angle. Its ‘shade’ visualizes the contours of the objects, whose organic remainders themselves have to be carefully macerated away from the metal copy. As a condition of its representation in this way, the original probe has to be eliminated altogether. The plane where instrument and object come in contact, the intersection point itself, is transformed into a new object endowed with resistance and permanence.
13
Nicolas Rasmussen. Picture Control. Stanford: Stanford University Press, 1997; Bruno Strasser. La fabrique d’une nouvelle science. La biologie moléculaire à l’âge atomique (1945-1964). Florence: Olschki, 2006.
Intersections
15
Fig. 4: Freeze-dried and tungsten-shadowed ribosomes from Escherichia coli.
8. Radioactive Isotopes and the Scintillation Machine In a review of the arsenal of instruments that became crucial for twentieth-century molecular biology, the technology of radioactive labeling occupies an important place. It represents an instrument of a particular character, and it has revolutionized the analysis of metabolic processes since World War II. Only now, however, have historians of the life sciences started to engage with this technology in detail.14 Radioactive tracing was introduced on a grand scale when radioactive variants of the most important atomic constituents of biomolecules became available in high yield as by-products of the nuclear reactors of the Manhattan project. In particular, these were isotopes of hydrogen, carbon, sulphur, and phosphorus. Immediately after the war, the American National Laboratory at Oak Ridge became the center for their production and for a distribution campaign known under the slogan of “atoms for peace.” Quickly distributed throughout the laboratory world of biological-chemi14
Cf. the special edition of the Journal of the History of Biology 39.4 (2006) on this topic with articles by Soraya de Chadarevian, Angela Creager, Karen Rader, and Maria Jesus Santesmases.
16
Hans-Jörg Rheinberger
cal and biomedical research, the practice of radioactive tracing consisted of replacing individual atoms in biomolecules with their radioactive isotopes and subsequently registering the decay events they produced. In principle, they behaved like probes that triggered tracer fires in particular metabolic reactions. This technology had a lasting influence on the molecular sciences. It allowed quantification in a range of concentration that was not accessible to classical chemical measurement. With the penetrating power of this technology, the registration range of signals became enhanced by about six orders of magnitude, from the micromolar to the picomolar. The scope of this dramatic turning point in the history of molecular biology may have escaped historical assessment because the new tool was of a distributive order and its functioning developed not as large scale imposing machinery but rather in a capillary fashion. Within two decades, between 1945 and 1965, it filtered through the biomolecular and biomedical sciences like a plexus of anastomoses. Radioactive markers can be seen as molecular instruments. They become deeply immersed in specific pathways of the metabolism, and there they sparkle and leave traces. They make possible a completely new form of biological chemistry. A reaction or a molecular component that one wishes to analyze no longer has to be isolated or purified before it can be measured. It can be visualized in vanishingly small concentrations and in a whole mixture of compounds. Individual reactions can be represented with high selectivity before a background full of noise. The autoradiogram makes visible the specific place of reaction right in the tissue or cell itself, that is, in situ. But the markers also allow one to follow molecular reactions in solution in the test tube. They are tags incorporated into the molecules whose movement one observes, and they do not alter their chemical constitution. Here, the intersection of the instrument with the epistemic object coincides with the epistemic thing whose traces, in contrast to the case of electron microscopy, can be followed on the wet side, without having to be transferred to the banks of the dry. In an exemplary fashion, radiolabeling exposes the deep paradox of the generation of traces. The creation of the trace goes hand in hand with the destruction of the isotope. At the very moment of the creation of the trace – and this irrevocably – its source decays. Consequently, the radiogram makes visible something that no longer exists at the place where the trace testifies to its presence, and where it now stands in for its past. The radiogram therefore is an instantiation par excellence of what makes a trace a trace: the absence of a reference.
Intersections
17
However, in order to get such records, the radioactivity of the probes has to be measured. In the case of the autoradiogram, a sensitive photo plate will do. Samples of another aggregation are more difficult to register. In parallel to the massive use of radioactive carbon, hydrogen, and sulfur, a new counter was developed that served as an alternative to traditional Geiger counter tubes and worked particularly well in the range of the weak ß-rays of these isotopes.15 The procedure was based on the transformation of the radioactive decay events into flashes of light in a liquid medium. The flashes were then sent through photomultipliers. Within a decade, the liquid scintillation counter conquered the laboratories of molecular biology, and together with the ultracentrifuge, became an emblem of cutting edge laboratory technology. The development of an automated counting device with a capacity of hundreds of samples had much more than a quantitative influence on molecular biological experimentation. As well as opening the possibility of serial testing, it allowed the development of qualitatively new experimental designs. The liquid scintillation counter offers a good example with which to study the effects of the introduction of a new instrument into an experimental system. Such instruments can endow the system with new qualities, although in this case we are dealing with a comparatively straightforward procedure: the simple introduction of a new counting device. The change in measurement plays out the potentials of a modified form of intersection between device and sample. On the long passage of the radioactive specimen out of the test tube and into the apparatus, the fluid intersection between the device and the sample reflects and complies with the inherent disposition to ‘wet’ experimentation in biochemistry and molecular biology. The liquid scintillation counter reconfigured this boundary in an extraordinarily flexible and versatile fashion. Despite its massive lead chamber and the electronic environment of the sample detector, the sample to be measured remained in the liquid environment of a glass or plastic vial (fig. 5).
15
Hans-Jörg Rheinberger. “Putting Isotopes to Work. Liquid Scintillation Counters, 1950-1970.” Instrumentation Between Science, State, and Industry (= Sociology of the Sciences Yearbook). Ed. Bernward Joerges and Terry Shinn. Dordrecht: Kluwer, 2001. 143-74.
18
Hans-Jörg Rheinberger
Fig. 5: First automated Liquid Scintillation Detector, steel shielding, dual elevators, 100 vial samples in four circular rows (1957).
9. Concluding Remarks By pushing the frontiers of analysis into the space of physics and chemistry, molecular biologists created a science that rests on a consistent ‘extracellular’ project. It relied on a battery of research technologies that brought forward a multiplicity of diverse intersections between the central epistemic objects of molecular biology, the biological macromolecules, and these new technologies. It made them measurable. A few instruments crucial for this endeavor have been described in this essay. Ironically, the consistent pursuit of this program, as indicated in the last section, led to a situation in which the former objects of epistemic interest themselves, the macromolecules, became transformed into an arsenal of molecular tools. Within the past 30 years, they have swallowed – as one might put it – the extracellular technology and transformed it, within the space of the cell itself, into the project of gene technology. The genetic engineers of today no longer construct the technology around the organism and adapt it to its surfaces, but insert their molecular instruments into the depth of the cell and let them act from within. They no longer analyze the organism; they recompose and reshape it. They are thoroughly constructive and synthetic. Under this kind of synthesis, the organism itself
Intersections
19
is being transformed into an instrument, not only of research, as it was already the case with model organisms, but of a cultural project at large. The intersection between nature and culture appears thus to have been reversed. Culture is now at work within the innermost core of nature. WORKS CITED Bachelard, Gaston. Le rationalisme appliqué. Paris: P.U.F., 1998. Chadarevian, Soraya de. Designs for Life. Cambridge: Cambridge University Press, 2002. Creager, Angela N.H. The Life of a Virus. Tobacco Mosaic Virus as an Experimental Model. Chicago: University of Chicago Press, 2002. Dierig, Sven. Wissenschaft in der Maschinenstadt. Emil Du Bois-Reymond und seine Laboratorien in Berlin. Göttingen: Wallenstein, 2006. Geimer, Peter, ed. Untot. Verhältnisse von Leben und Leblosigkeit. Preprint No. 250. Berlin: Max-Planck-Institut für Wissenschaftsgeschichte, 2002. Geimer, Peter. “Was ist kein Bild? Zur ‘Störung der Verweisung.’” Ordnungen der Sichtbarkeit. Fotografie in Wissenschaft, Technologie und Kunst. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 313-41. Kohler, Robert E. Lords of the Fly. Chicago: University of Chicago Press, 1994. Mayer, Andreas. “Autographien des Ganges. Repräsentation und Redressement bewegter Körper im neunzehnten Jahrhundert.” Kunstmaschinen. Spielräume des Sehens zwischen Wissenschaft und Ästhetik. Ed. idem and Alexandre Métraux. Frankfurt a.M.: Fischer, 2005. 101-138. “Radiobiology in the Atomic Age: Changing Research Practices and Policies in Comparative Perspective”. Journal of the History of Biology 39.4 (2006). Rasmussen, Nicolas. Picture Control. Stanford: Stanford University Press, 1997. Rheinberger, Hans-Jörg. “Vom Mikrosom zum Ribosom. ‘Strategien’ der ‘Repräsentation’ 1935-1955.” Die Experimentalisierung des Lebens. Ed. idem and Michael Hagner. Berlin: Akademie Verlag, 1993. 162-87. Rheinberger, Hans-Jörg. “Putting Isotopes to Work. Liquid Scintillation Counters, 1950-1970.” Instrumentation Between Science, State, and Industry (= Sociology of the Sciences Yearbook). Ed. Bernward Joerges and Terry Shinn. Dordrecht: Kluwer, 2001. 143-74. Rheinberger, Hans-Jörg. “Präparate – ‘Bilder’ ihrer selbst. Eine bildtheoretische Glosse.” Oberflächen der Theorie (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1,2). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2003. 9-19. Schickore, Jutta. “Fixierung mikroskopischer Beobachtungen. Zeichnung, Dauerpräparat, Mikrofotografie.” Ordnungen der Sichtbarkeit. Fotografie in Wissenschaft, Kunst und Technologie. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 285-310. Schickore, Jutta. “(Ab)using the Past for Present Purposes. Exposing Contextual and Trans-Contextual Features of Error.” Perspectives on Science 10 (2002): 433-56. Schleiden, Matthias Jacob. Die Botanik als inductive Wissenschaft behandelt. Zweite, ganzlich umgearbeitete Auflage der Grundzüge der Wissenschaftlichen Botanik. Zweiter Theil. Morphologie, Organologie. Leipzig: Engelmann, 1846. Strasser, Bruno. La fabrique d’une nouvelle science. La biologie moléculaire à l’âge atomique (1945-1964). Florence: Olschki, 2006. Virtual Laboratory for Physiology. Online: http://vlp.mpiwg-berlin.mpg.de
DIETER MERSCH
Representation and Distortion: On the Construction of Rationality and Irrationality in Early Modern Modes of Representation Introduction In Proclus’ commentary on Euclid’s Elements, we find the remarkable statement that Pythagoras “gave the geometrical science a new form, that of a free discipline.”1 And in his work On Arithmetic, Aristoxenus added that he also “freed dealing with numbers from its practical uses . . . and explained things as being a picture of numbers. For a number also contains everything else, and between all numbers there is a reciprocal rational relationship.”2 This understanding of mathematics as a pure science (mathémata) became dominant, removed from all direct uses and finding its sole purpose in the formulation of general rules and the demonstration of their validity. Having become autonomous, it was possible for mathematics to create a number of laws and principles that derived exclusively from an internal game of rationality and not the observation of nature. The Pythagorean rule was certainly the most famous of these, but not the most fundamental. Pythagoras had, rather, attempted to establish a cosmological order on the basis of the principle of the number and its derivations, and which found its internal fulfilment and its secret in the figure of the Tetraktys. This system remained incomplete, however, as it was precisely this Pythagorean rule that led to the scandal that unsettled Greek mathematics as much as the discovery of logical paradoxes had, namely the problem of the “incommensurable”: the impossibility of representing the side and the diagonal of a square in the same measurement system. At this point a moment of irrationality broke into the integer, i.e. rational proportions – a second secret that the Pythagoreans kept as safe as the 1 2
Jaap Mansfeld, ed. Die Vorsokratiker I. Stuttgart: Reclam, 1987. 169. Ibid. 169.
Representation and Distortion
21
Tetraktys. In his description of the life of Pythagoras, Iamblichus reports that this secret was guarded fiercely to the extent that those who broke the taboo were adamantly excluded from the community, indeed a gravestone was set up for them, “as if the former friend had completely departed from human life.”3 Every formal system or calculation creates its own limits where it breaks down. Such limits within a system mostly take aporetic forms inasmuch as they are demonstrable but cannot be solved or integrated into the system. This holds true for the Tetraktys because, in accordance with its own structure and entitlement, it allowed the creation of quantities that were not plausible within its own terms of reference. An analogous phenomenon can be seen in the mathematical thinking of the early modern era, especially in the predominant projection geometry. This paper is an investigation of only one detail of geometrical thinking that played an important role in the development of art and visual representation from the sixteenth to the eighteenth centuries: the dialectic of central perspective and anamorphosis. The central thesis is that anamorphosis held the same status in relation to perspective constructions as the incommensurable did in relation to the system of rational numbers or as the paradox did in relation to the method of syllogisms in Greek mathematics and logic. For at the same time that a system of vanishing points created the illusion of spatiality on a flat surface, it also created distortions that undermined this representation of reality once more. In particular, anamorphotic representation is shown to be an element of the construction of perspective itself which connects it to the problems of the incommensurable and the paradox: in the repertoire of projective geometry, something that blurs the projection and destroys the figure can be produced. 1. Central Perspective Ever since the fifteenth century the emancipation of the visual arts has been supported by mathematics in the shape of geometry, especially the theory of the five Platonic bodies, the laws of proportion and the methods of constructing perspective.4 Together these things formed the basis 3 4
Ibid. 171. For a discussion of the role of “Platonic bodies” in the sixteenth century, cf. Hans Holländer. “Mathematisch-mechanische Capriccios.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 352. Dürer saw it as the quintessence of divine order. The doctrine can be seen in the illustrations that
22
Dieter Mersch
of optics, picture theory and celestial mechanics, as can be seen in Leonardo da Vinci’s illustrations for the mathematical investigations of Luca Paciolo5 as well as in Johannes Kepler’s Mysterium Cosmographicum of 1596. The structure of the image was in no way limited to the projection theory of central perspective: This was rather only one method of rational representation from a fixed point of view, albeit the most important one. In fact, the central perspective made the systematic connection between the subjectivity of the view and the constructivity of mathematics possible. Two things are included here: i) firstly, the relativisation of the view in the imaginary position of the observing or constructed subject, and ii) secondly, the identification of subjectivism and constructivism in a way that can most clearly be seen later on in terms of philosophy in the work of Immanuel Kant. As is well known, this artistic project goes back to Filippo Brunelleschi, whose picture experiments have been lost. In his account of the life of the artist, Antonio Manetti writes about two panels made between 1418 and 1425, mounted in the baptistery in Florence, which presented an almost perfect play of illusion. One could contemplate this through a hole drilled in a canvas screen that captured the gaze, a trompe-l’œil effect being achieved thanks to a sophisticated system of mirrors. This arrangement seems a perfect example of the early connection between the eye, mathematics and mechanics: for in order to enable the exact perception of pictures a fixed point of view had to be chosen so that the vanishing point, visual axis and the standpoint of the viewer related to each other. This demanded a good deal of calculation because only from this point could a three dimensional impression be created – with a picture measuring 120 x 120 cm at a distance of a little more than 2.5 metres.6 We are dealing with the geometrical construction of a percep-
5
6
Leonardo made for Luca Paciolo’s mathematical experiments. Just as important, especially for the construction of the body in space, was the science of proportion, to which Paciolo dedicated his book, De Divina Proportione (1509), in which proportion and perspective are dealt with as separate areas despite the many things they have in common. Together these things formed the basis of picture theory and art, as well as optics in general and celestial mechanics. For example, Kepler’s Mysterium Cosmographicum (1596) is based on them, with Platonic bodies organising the cosmos. The theory of proportion, with which Paciolo dealt in De Divina Proportione (1509), was equally important for the construction of bodies in space, whereby the theory of proportion and the perspective are treated as different fields of mathematics despite their various intersections. Leonardo da Vinci recommended a distance of three times the height of the object concerned.
Representation and Distortion
23
tion effect whose apparent purpose was to fix the eye of the artist and that of the observer on the same clearly calculable point. This required apparatuses that were usable on both parts in principle:7 on the part of the artist as a perspective apparatus, as one can see in the illustrations of Leonardo and Albrecht Dürer, or on the part of the viewer as “Perspectyfkas,” as was prevalent later, especially among the Delft School – a box-like structure that brings the view and picture into the right position in order to produce the appropriate trompe-l’œil effects in a controlled way. Perspective vision, which first of all had to be ‘instrumentalized’ and learned in this way, meant that the viewer fixed his eyes onto an established point in a way that could be compared to conditioning, which, controlled by the trained eye, gave the beholder a synopsis of the world thus presented. It required the replacement of traditional perception by a “stereometrical time-space construct” that dominated the view.8 Daniele Barbaro, a friend of Andrea Palladio, wrote in his Pratica della Prospettiva of 1569: I give instructions on how the rays of sight . . . have to be directed from one specific marked point, rays which match the lines of nature. Whatever scenes you might see . . . it is certain, that this whole practice is summed up in three terms and their understanding, namely eye, rays of sight and distance.9
“This system, wretchedly quarried out of the cosmos and history,” Rudolf zur Lippe adds in his study of the geometricalization of man in the early modern era, “describes precisely the circumstances of the concept of perspective, where there is only one standpoint and one presentation of the corporeal world based around it,”10 as it appeared in a perspective box, for example. All representations were marked with this constructedness that obeyed simple rules. This involved a mystification of the rigid and ‘possitioned’ observers, who, as Erwin Panofsky later put it, created a “systematic space” in which an eye encounters an equally motionless object.
7 8
9 10
Leonardo argued against this. Cf. André Chastel, ed. Leonardo da Vinci. Sämtliche Gemälde und die Schriften zur Malerei. Munich: Schirmer/Mosel, 1990. 255. Cf. Gottfried Boehm. “Zwischen Auge und Hand. Bilder als Instrumente der Erkenntnis.” Mit dem Auge denken. Strategien der Sichtbarmachung in wissenschaftlichen und virtuellen Welten. Ed. Bettina Heintz and Jörg Huber. Zurich and New York: Spinger, 2001. 49f. Quoted in Rudolf zur Lippe. Vom Leib zum Körper. Naturbeherrschung am Menschen in der Renaissance. Reinbek bei Hamburg: Rowohlt, 1988. 266. Ibid. 200f.
24
Dieter Mersch The two-dimensional realisation of three dimensional space, as can be seen in the art of the Renaissance, with its measured clarity and the consistent shaping of all parts, the common vanishing point of all parallel lines and the unified mode of determination of distance . . . is a bold abstraction,
wrote Arnold Hauser. “Central perspective produces a mathematically correct space, but not a psychologically real one . . .”11 The theory, technique and art of perspective therefore have all the signs of a dispositive; that is, a general scheme that the view has to fit into. As such it is not passive, but productive. In other words, central perspective is itself an instrument of the production of visibility; it does not obey the visio, the view, but creates visuality. The organization of the visible based on the logic of proportional measurements and relations, as was carried over into other arts such as drama, dance and even fencing, was also affected.12 Alongside the mathematics of projection geometry, which was developed in a more experimental and intuitive rather than systematic way by Leon Battista Alberti, Leonardo and Dürer (Leonardo’s notes on “linear perspective” are especially enlightening in this context),13 and whose actual formulation was left to the “painting mathematicians” of the seventeenth century, the unified basis was the control of space by means of technology. Together with the rationalisation of perception and the principles of aesthetics, it formed the basis of the compulsory regulae of art, through which it hoped to secure 11
12
13
Arnold Hauser. Sozialgeschichte der Kunst und Literatur. Munich: C.H. Beck, 1969. 357. For a discussion of the constructedness of central perspective, cf. Erwin Panofsky. “Die Perspektive als ‘symbolische Form.’” Vorträge der Bibliothek Warburg 25 (1924), reprinted in H. Oberer and E. Verheyen, eds. Aufsätze und Grundfragen der Kunstwissenschaft. Berlin: Spiess, 1992. 258-330; Nelson Goodman. Languages of Art. Indianapolis: Hacket, 1976. 10ff.; John Berger. Ways of Seeing. London: BBC/Penguin, 1972. 16: “The convention of perspective, which is unique to European art and which was first established in the early Renaissance, centres everything on the eye of the beholder.” This is also true in terms of perception psychology, since in principle pictures in perspective and real objects do indeed throw the same or similar images onto the retina. However, this is only true for the abstract perception of a non-moving eye, which does not exist. In fact, things are observed by a constantly moving eye. “For this reason, monocular seeing with a moving eye is not central perspective, but rather a synthesis of a variety of individual central perspectives from different centres.” Wolfgang Haack. Darstellende Geometrie (= Sammlung Göschen, vol. 144). Berlin: de Gruyter, 1969. 7. “The whole development of art is an integral part of the great general process of rationalisation . . . and as central perspective is only the mathematicisation of space, and proportion only the systematization of individual forms in representation, so all criteria of artistic quality are gradually subject to rationality and all rules of art are rationalised.” Hauser. Sozialgeschichte der Kunst und Literatur. 293. Quoted in Chastel. Leonardo. 245ff.
Representation and Distortion
25
its position within the septem artes liberales, in which geometry held first position. In this way the visual arts were lifted out of the realm of the mere artes mechanicae, which were seen as nothing more than an aid, and made an independent scientia. This happened around the time of Jan Vermeer, and his painting “The Art of Painting” bears eloquent witness to this phenomenon.14 Two concepts intersect here: firstly, the mathematical concept of geometricalization through the construction of vanishing points, intersecting lines and metricisation of space as a prerequisite for a framework of equal squares that was used by Alberti and which Leonardo and Dürer replaced with mechanical instruments, and secondly, the aesthetic concept of proportion and symmetry that proclaimed the principles of “divine” harmony and beauty, as handed down and continued by the ideals of Classical antiquity. Only the latter distinguished the artist from the scientist and made art border on magic because it showed the regularity of all beauty as well as its calculability and construction. This also meant that mathematics and aesthetics showed themselves to be closely intertwined. Mathematics dealt with what could be constructed with a pair of compasses, a ruler and the basics of Euclidian geometry, whereas aesthetics dealt with the stable morphology of nature. The former followed the Classical episteme, which was passed on to the Middle Ages and canonised in the medieval Quadrivium, the latter followed mimesis, which was also based on mathematical, that is rational, foundations.15 2. Conic Section Theory The systematic connection between art and science in the sense of Classical ars, as was still taken for granted in the sixteenth century, becomes clear. The distinction between the two can only be seen to start in the eighteenth century with a new ideal of objectivity.16 It is also 14
15
16
Cf. Dieter Mersch. “Ästhetischer Augenblick und Gedächtnis der Kunst. Überlegungen zum Verhältnis von Zeit und Bild.” Die Medien der Künste. Beiträge zur Theorie des Darstellens. Ed. Dieter Mersch. Munich: Fink, 2003, especially 152ff. Cf. also Sybille Krämer. “Zentralperspektive, Kalkül, Virtuelle Realität: Sieben Thesen über die Weltbildimplikationen symbolischer Formen.” Medien-Welten, Wirklichkeiten. Ed. Gianni Vattimo and Wolfgang Welsch. Munich: Fink, 1998. 27-37. Cf. also Lorraine Daston and Peter Galison. “Das Bild der Objektivität.” Ordnungen der Sichtbarkeit. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 29-99, and Martin Kemp. Visualizations. The Nature Book of Art and Science. Berkeley: University of California Press, 2001.
26
Dieter Mersch
clear that at first, and indeed already in the early Renaissance, art foreshadowed mathematics, and therefore its research and achievements in relation to the construction of perspective formed the paradigm which later, especially in the seventeenth century, was unified by mathematicians and opticians, turned into an axiom and canonised as a technique. On the other hand, from the very beginning experimental artists in particular looked for the basic principles in mathematics, which gave the still rudimentary study of perspective the dignity of a revelation of the divine order. Art that used this kind of geometry in order to establish itself as a science created visible miracles that combined rationality and mysticism so that the rationalism of mathematics did not lead to the destruction of the theological order, but confirmed it in the form of an idealized divine architecture. By penetrating universal laws, man read the book of nature, imitated God and acquired an insight into the eternal construction of His creation, thereby taking on the mantle of an alter deus – a concept that later became the idea of the genius and its worldly counterpart, the engineer, who became the heroes of the nineteenth and twentieth centuries. The painting and art of the fifteenth and sixteenth centuries therefore cast a shadow. Their mathematical investigations, which were still oriented towards empiricism and practical requirements, anticipated the calculisation of nature that found its theoretical (that is particularly algebraic) basis about one hundred years later with the work of René Descartes and later Gottfried Wilhelm Leibniz.17 Only subsequently did the gradual conversion of ars as a mathematical and miraculous art into an equally methodical and scientific techné take place, experiencing its heyday as a technical science and then as a technoscience in the nineteenth and twentieth centuries. By speaking of perspective as a “dispositive” of the construction of visibility, one is dealing with a complex dialectic between art and science. If painting anticipated mathematics in practical procedures, then mathematics gave art the decisive impulse towards autonomy. Although the basic ideas of projection geometry are already present in the first volume of Euclid’s Elements – as images using perspective found in Pompeii show – the essential contribution of early modern art was the systematization of its principles that allowed the development of representational geometry as a self-contained mathematical discipline, independent of art. In fact, one sees a foreshadowing of this in Alberti’s 17
Cf. also Sybille Krämer. “Kalküle als Repräsentation. Zur Genese des operativen Symbolismus in der Neuzeit.” Räume des Wissens. Repräsentation, Codierung, Spur. Ed. Hans-Jörg Rheinberger, Michael Hagner, and Bettina Wahrig-Schmidt. Berlin: Akademie Verlag, 1997. 111-22.
Representation and Distortion
27
treatise Della Pittura of 1435, in which he defines a picture as a vertical cut through the “visual pyramid.”18 The top of the pyramid represents the viewpoint of the viewer, whose imaginary continuation is the vanishing point on the picture, which is only possible within considerable mathematical restrictions. On this theme Leonardo wrote, Perspective is nothing more than seeing a scene from behind a flat and transparent piece of glass that has all the objects beyond the glass drawn on it. They are led through pyramids to the point of the eye and the pyramids are cut from the mentioned piece of glass.19
However, Alberti and Leonardo proceeded heuristically and it was only in Piero della Francesca’s De Prospectiva pingendi (ca. 1474) that a systematic description of the mathematical basis of perspective could be found. Almost one hundred years later, Palladio, in his Quattro libri dell’Architettura of 1570, rendered this more precisely, creating a system that endured until the eighteenth century and that indeed is still the accepted basis for the logic of images in science and technology and the virtual visualizations of computer graphics. However, it was the mathematicians of the seventeenth century, especially Jean-François Nicéron, who took general conic section theory as the starting point of their geometry, thereby made all the different methods of constructing perspective (with a rising line of refinements and perfections, from all points and angles, from various heights and distances, with slanting axonometry or the projection of various vanishing points, etc.) possible and who explained these ways as leading back to a uniform model.20 Since then, in principle, every image and every figure has been subject to the theory of the conic section – just as contrariwise the 18 19 20
Leon Battista Alberti. On Painting. New York: Penguin, 1991. 55f. Alberti’s starting point is the picture. In fact it is a visual cone. Quoted in Chastel. Leonardo. 246. Cf. Holländer. “Mathematisch-mechanische Capriccios.” 353f. and 340f. He continues: “With him, perspective means central perspective as understood by Alberti and Dürer, who are quoted along with a dozen other authors . . . The concept of central perspective that is still valid today is only an elementary prelude for him, albeit a prerequisite for that which followed. He takes a generalized concept of central perspective, which no longer has much in common with that of Alberti for granted. He does not speak about a visual pyramid that relates to the square nature of the board, but relates to the visual cone. In specific Hans Holländer. “Anamorphotische Perspektiven und Cartesianische Ornamente. Zu einigen Gemälden von Jean-François Nicéron.” Rezeption und Produktion zwischen 1570 und 1730. Festschrift für Günther Weydt. Ed. Wolfdietrich Rasch, Hans Geulen, and Klaus Haberkamm. Munich and Bern: Francke, 1972. 59. Holländer points out that before Nicéron, the creation of perspective images was described as a special case of the science of conic sections in optics by Franciscus Aguilonius in 1613.
Dieter Mersch
28
term central perspective covers all types of construction, even the distorted ones that would not strictly speaking fit into the system. Johann Heinrich Lambert, who was responsible for many important steps in the discovery of non-Euclidian geometry, related the study of perspective to the art of seeing, which had previously been known as Optica. At the start of the seventeenth century, Perspectiva and Optica meant the same thing.21 They became the norm in terms of seeing, a norm which included the unity of three systems: i) the pictorial geometry of painting, ii) conic section theory and its justification in the analytical geometry of Descartes, and iii) optics in the physical and technical theory of lenses. This means, on the other hand, that the theory of conic section and the theory of sight merged and defined the totality of the visible. The talk of a “dispositive of central perspective” is precise and legitimate in this context: seeing in perspective became a condition for the possibility of seeing and thereby became a constitutive principle of the visible in general: at the same time it denoted its “transcendental” character. However, the establishment of the geometry of perspective as an independent discipline was connected with a strange shift that sought to separate the mathematical rules from the logic of representation and create a pure science of conic sections, which dealt not only with the circle which is so important for art, but also with ellipses, parabolas and hyperbolas. When this is transferred back to painting, the picture as a cross-section through the visual cone can be described in all its distorsions as was demanded by Alberti – particularly the slants that are characteristic of “anamorphotic projections.” In other words, the dispositive reveals its disruptions and puzzles at its margins. It is in these strange border zones that its system displodes. Even as it was discovered and systematically studied by the painters of the fifteenth and sixteenth centuries in order to solve the problem of distortion in large wall, ceiling or vault frescos, these border zones shook its mathematical foundations. Thus Leonardo distinguished between several different kinds of perspective in his writings on the Problems of Painting, namely between “natural” and “accidental”22 or “artificial,” which included ball-shaped and mixed constructions like the anamorphosis, that he called “monstrous.”23 Leonardo’s experiments are well known – they are the oldest known anamorphoses in the form of studio objects that demonstrate 21 22 23
Cf. Holländer. “Mathematisch-mechanische Capriccios.” 342. Quoted in Chastel. Leonardo. 254f. Ibid. 257.
Representation and Distortion
29
that they are to be understood as part of a system and of the mathematics of central perspective without having a readily recognizable representation. They remain within their frames in order to break them open and thereby show the limits of the system. At the same time they undermine its internal rationality by producing figures within its order that seem to mock reason. 3. Anamorphosis One could describe anamorphosis as a kind of “Wunderkammer of perspective construction.” Born out of this context, it dissolves representation and blurs the lines of visibility. It does not, however, deal with the representation of the unrepresentable, as the art of Romanticism would, but rather with making a limit known, a limit which as such requires the whole dispositive and the mathematical system of Euclidian geometry and its instrumentation and confirms it in its validity. However, at the points where this system breaks down, mysteries and irrationalities creep in. That is why Leonardo writes of monstrosities. The word “anamorphosis” describes this phenomenon. It refers to morphé, the shape or form whose distortion or deformation is expressed in the prefix ana: anamorphosos, which emerged as a neologism in the baroque period, means the “deformed” or the “unformed” in this context. This origin can be traced back to the Magia Universalis by the mathematician Caspar Schott in the early seventeenth century, which dedicates one chapter of seventy pages to the subject of Magia Anamorphotica and sees it as a pictorial miracle because it reveals what is hidden from normal perception: a face, a scene or a landscape which emerges from a confusingly unclear image when seen from an extreme angle of nearly 180°.24 This does not mean anamorphosis is obscure or chaotic, but rather that the distortion has a regular pattern that worked like a false mirror of rising rationalism. That is, anamorphosis marks a fissure that breaks open within the interior of the representation itself. It is related to the logic of central perspective in much the same way as the capriccios of the arts mechanicae are to the useful apparatuses of the baroque. They were like a rational means of generating the irrational – this is why they were, especially later, kept in curiosity cabinets with talking dolls, writing machines and artificial chess players. The anamorphoses did not 24
Cf. Thomas Eser. Schiefe Bilder. Die Zimmernsche Anamorphose und andere Augenspiele aus den Sammlungen des Germanischen Nationalmuseums. Nuremberg: Germanisches Nationalmuseum, 1998. 7.
Dieter Mersch
30
however leave the narrow realm of perspective. As a riddle which complied with the conditions of rationalism as much as it anulled them, it rather served calculated concealment, the cryptic and the enigmatic. From the very beginning therefore the history of anamorphosis was maintained – as a subversive inscription within the system of perspective and its geometry – in this conflict between secret craftsmanship on the one hand – which it carefully maintained and guarded, especially in relation to the transmission of knowledge about the techniques of its production – and the representation of an open secret on the other hand, like a mask which both reveals and disguises. Both sides are therefore always present: the perception of the inexplicable to which the astonished observer surrendered, and the codification of a mystery within the meaning of an artificial secret code whose key was waiting to be discovered. One can therefore distinguish between two kinds of anamorphotic artistic production. Firstly, images with allegorical and religious content, for example Hans Holbein’s The Ambassadors of 1533, certainly the most famous painting with an anamorphotic skull in the foreground, or William Scrot’s portrait of 1584 with its disfigured face that appears charming if seen from the side. They are similar in their mode of production to an optical distortion, where one sees death – like beauty – only when one does not see life – or decay – and vice versa. The second kind includes works of a political or critical nature, such as the elaborate 1535 Zimmern anamorphosis which shows a complex pictorial riddle in the form of a chest with relief pictures showing the Graf von Zimmern on one side, and his wife Amalia von Leuchtenberg on the other. Other contemporary examples include the woodcuts of Erhard Schön, following Dürer, such as “It’s Over, You Old Fool,” also from 1535, or “What Do You See?” from 1538 – drastic representations of frivolous subjects which at the same time spread a moral message.25 In addition, one finds idyllic landscapes that anticipate the sufferings of Christ with His crown of thorns, secret images of Ludwig III praying in front of a crucifix or of Saint Francis in ecstasy. The largest anamorphosis ever completed is by Emmanuel Maignan in the monastery of San Trinità dei Monti in Rome, which literally has a second meaning alongside what is represented in the foreground. Historically it is possible to divide this phenomenon into three rough periods: i) a period that stretches from the late fifteenth to the middle of the sixteenth century and which had its apogee in the Augsburg-Nuremberg school around Dürer, with Schön and the Zimmern anamorphosis; 25
Ibid. 7ff.; also Holländer. “Anamorphotische Perspektiven.” 54.
Representation and Distortion
31
ii) the theoretical phase between 1620 and 1670 which systematically strove to describe the distortions of perspective and to explain them mathematically and geometrically. At the heart of this phase is Nicéron’s comprehensive book, La Perspective curieuse ou Magie artificielle des effets merveilleux (1638), which was republished under the title Thaumaturgus Opticus, after the author’s early death, by the Minorite priest Marin Mersenne, a friend of Descartes.26 This book contains nothing less than an attempt to create a complete classification of all perspective-based forms of representation on the basis of the conic section theory alongside other forms such as conical or catoptrical anamorphoses on conical, spherical and cylindrical surfaces. It is worth observing that it is this period that saw the creation of the first automated machine for producing anamorphoses, particularly the cylindrical mirror anamorphosis, such as the one designed by Jacob Leupold from Leipzig. At the turn of the seventeenth and eighteenth centuries he produced whole series of engravings of this kind, which he placed next to corresponding anti-distortion mirrors.27 This era can be distinguished from iii) the third period of mannerism in which anamorphosis is reduced to nothing more than an optical skill, curiosity or game, appearing in the Wunderkammer of the eighteenth century with the purpose of entertaining the spectator, until its trail is lost. Although it continued to be popular and indeed became widespread, it had become a showpiece used for theatrical effect. For this reason, the nineteenth century scarcely showed any interest in it, which was related to the fact that it was no longer an example of a dialectic of rationality and irrationality, but simply a question of the Other of reason in the form of the limits of the visible, the nonrepresentable or the representation of the unrepresentable. Nevertheless it seems clear that from its very beginning anamorphosis formed a mode of reflection within the mathematics of perspective, as the examples of Leonardo, Michaelangelo Buonarroti and Dürer (who must also be counted among the initiators) show.28 At first it was explored as an experimental method of balancing out extreme angles of vision, but as early as the sixteenth century methodical instructions such as Daniele Barbaro’s 1569 Practica della Prospettiva followed, which used perspective apparatuses to divide the picture up into squares in order to transfer the strange images, anamorphotic extremes and cones 26 27 28
Holländer. “Anamorphotische Perspektiven.” 55 (Note 11) and 60. Cf. Eser. Schiefe Bilder. 118. In this sense, one could interpret the self-portrait by Michelanglo on the skin of Bartholomew in the Sistine Chapel as an early anamorphosis. Cf. Holländer. “Anamorphotische Perspektiven.” 53.
32
Dieter Mersch
and cylindrical reflections exactly. They rationalized the phenomenon in the same way as the often broken reflection systems of the baroque which sought to make the non-visible visible and the hidden representable. However, one can only speak of a “real anamorphotic perspective,” as Jurgis Baltrušaitis noted in his compendium on the phenomenon of anamorphosis, “when there is a break between the form and its recording.”29 Thus anamorphosis nestles like a parasite in the dispositive of visibility in order to hollow out this order. 4. Angled Conic Sections and the Beginning of Non-Euclidian Geometry As a result of this, anamorphotic distortion within the geometry of central perspective turns out to be an immanent possibility which does not show the object but negates it. Its construction remains strictly within the rules of perspective, in such a way that it breaks up the perspective of its representation at the same moment that it uses it. It therefore belongs in the category of optical paradoxes: one could speak of a perspectival construction of aperspectivity. The supposed optical miracle that contemporaries saw in anamorphotic distorsion is itself part of the system and represented by it in order to destroy the representation. As a break away from the figure, it remains in the realm of the figurative to play out the rational dispositive in all its variety, to exaggerate it until the figure, the representation breaks down. In this sense, anamorphosis is similar to the method of the aporetic. Whereas the mathematics of central perspective attempted to represent the real on a flat surface in a way that appeared spatial, anamorphosis in contrast allowed representations which thwarted any attempt at a spatial reading. In this way anamorphosis remained within reason as a kind of “poetic abstraction,” which Baltrušaitis describes as a “great unknown,”30 that is constantly countered by new opposition and fascinated Jean Cocteau as much as Jacques Lacan or Roland Barthes. If one takes a closer look at its mathematical format, however, anamorphosis’s otherness has geometrical rules that regulated the structure 29
30
Introduction by Jurgis Baltrušaitis to Freed Leeman, Joost Elffers, and Mike Schuyt. Anamorphosen. Ein Spiel mit der Wahrnehmung, dem Schein und der Wirklichkeit. [Exhibit. cat.] Cologne: DuMont, 1975. 6. Cf. Jurgis Baltrušaitis. Anamorphoses ou perspectives curieuses. Paris: Perrin, 1955; expanded edition: Anamorphoses ou Magie artificielle des effets merveilleux. Paris: Perrin, 2nd ed. 1969; new edition: Anamorphoses ou Thaumaturgus opticus. Paris: Flammarion, 1996. 7. The use of the title and its variations clearly follow Nicéron’s work on perspective of 1638.
Representation and Distortion
33
of space for over 2000 years. Because of their early axiomatisation these rules are known as “Euclidian.” This becomes particularly clear when one looks into the theory of anamorphosis that Nicéron presents in his Perspective curieuse. Characteristically Nicéron deals with anamorphosis as a part of conic section theory, and indeed, not as a level plane but as a slanted section at an acute angle to the visual axis, which had to be balanced correspondingly with the viewpoint of the observer – from the side.31 If the totality of the visible is completely defined by the theory of conic section, then anamorphosis is created as an extreme phenomenon by means of elliptical, parabolical or hyperbolical strechting. The geometrical rules are not broken, but alternated. That means that we are dealing with a shift towards a different kind of geometry. In the context of Euclidian space, anamorphosis conceals the discovery of a nonEuclidian understanding. Entirely describable within the Euclidian system, it nevertheless creates a picture that, especially when in the form of a cylindrical reflection, appears to be a Euclidian image of a non-Euclidian landscape. The shift would ultimately be complete if one introduced imaginary sizes instead of real ones, curved surfaces instead of flat ones or logarithmic scales instead of linear ones, as was available at the beginning of the seventeenth century. Nevertheless, it was a long and difficult process to break out of the chains of the Euclidian world view, which even Immanuel Kant still took for granted.32 From a mathematical point of view, anamorphosis is not an example of bizarre formlessness sprung from a chaotic spirit, but rather all sections are perceived as essentially equal through the cone, as the various geometries showed themselves to be isomorphic.33 There is therefore a Euclidian and a non-Euclidian con31
32
33
This is similar to when spherical surfaces are transferred onto a square framework with polar coordination. The distortion can be reversed by using a spherical mirror which is supplied with it. Cf. Hans Holländer. “Spielformen der Mathesis universalis.” Erkenntnis – Erfindung – Konstruktion. 341ff. No less a figure than Carl-Friedrich Gauß held back with his research into nonEuclidian geometry because of his fear of the “cries of the Boeotians,” as he wrote in his letter to Bessel from 27 January 1829, as its publication would clash with the ideas of his time. Cf. Oskar Becker. Grundlagen der Mathematik in geschichtlicher Entwicklung. Frankfurt a.M.: Suhrkamp, 1975. 179. The isomorphism of geometry is the content of Klein’s theory of transformation groups. Cf. ibid. 195ff., the chapter “Über die sogenannte Nicht-euklidische Geometrie.” In particular, Klein distinguished between elliptical, parabolic, and hyperbolic geometry, according to cone section theory, where Euclidian geometry corresponds to the parabolic. All propositions in hyperbolic geometry can be transferred to Euclidian geometry and vice versa. By reciprocal calculation between the
34
Dieter Mersch
struction in every image. The anamorphosis distorts the Euclidian system to its breaking point, but does not make it relative, and instead reveals something else as its point of origin. However, the crossing of limits, the deviation from the schema, would amount to leaving the system of classical Optica and its representational power, which the seventeenth century would not have considered. The seed that it created germinated in the nineteenth century with the various drafts of non-Euclidian geometry and came to fruition in the twentieth century with its axiomatisation and formalisation. Significantly this only became possible, however, when the limits of contemplation were left behind and it was thus possible to establish geometry as a pure construct and abstract model of space.34 The decisive turning point was a rigorous uncoupling of eye and rationality. At the beginning of the nineteenth century Bernhard Bolzano complained that many characteristics of geometrical objects were inferred directly from sight – instead of consistently leaving out all visual reference35 – but in contrast non-Euclidian spaces include structures which deny visual interpretation. More than fifty years after Bolzano, this fact caused Felix Klein to get rid of the “mathematically non-essential sensual image” and to concentrate on the “space of only one multitudinously extensive diversity.”36 In contrast to Classical geometry and the construction of perspective in the early modern period, from now on we are no longer dealing with the visible, but with the logical possibilities for which only their lack of contradiction counted, which destroyed the link between calculation and perception forever. Seen from the opposite point of view, that means that the possibility of non-Euclidian geometry did not follow from seeing and modelling its distortions, but from the logic of axiomatisation
34 35 36
geometries all proofs can be achieved algebraically and without recourse to perception. Cf. R. Baldus and F. Löbell. Nichteuklidische Geometrie (= Sammlung Göschen, vol. 970/970a). Berlin: de Gruyter, 1964. 143. Cf. Herbert Meschkowski, ed. Lust an der Erkenntnis. Moderne Mathematik. Munich and Zurich: Piper, 1991. 55. Felix Klein. “Erlanger Programm.” Quoted in Becker. Grundlagen der Mathematik. 198. The substitution of spatial categories with n-dimensional diversity goes back to Riemann, who at the same time replaced perceived space with abstract space. Cf. ibid. 185-93, the chapter “Über die Hypothesen, welcher der Geometrie zugrunde liegen.” A purely axiomatic (i.e., symbolically based) geometry was tackled by Moritz Pasch in 1882 in his lectures on the new geometry and by David Hilbert in his Foundations of Geometry. Cf. the corresponding extracts in Meschkowski. Lust an der Erkenntnis. 60ff., and Becker. Grundlagen der Mathematik. 203ff.
Representation and Distortion
35
(i.e. the transformation of its mathematics into pure syntax). The expression “discovery” which is used in this context and which can lead to misunderstandings, refers to this paradigm shift. Its mathematical existence is based entirely on its consistent constructability, with no pretence of dealing with reality. There is, therefore, no non-Euclidian geometry in the real world that could be made visible, as anamorphosis as its model suggested, instead “they exist” only in a logical world, whose symbolic construction allows no visual representation. Paradoxically, non-Euclidian geometry frees mathematical geometry from the visual nature that is its basis. Indeed, Euclidian structures are seen merely as a special case and as inferior to non-Euclidian structures because their points, lines, angles and surfaces are defined exclusively in terms of form. Nevertheless, the history of anamorphosis is part of the background to the rise of non-Euclidian geometry because it opened theoretical possibilities thanks to its artistic uses and its visual distortions, out of which in turn followed axiomatization. It literally tore down the usual conceptions of space. It is therefore less a description of a minor field, a footnote to the history of art of the sixteenth and seventeenth centuries, than a demonstration that the dynamic of a culture cannot be written in a one-sided way, either from the viewpoint of technology, or from the viewpoint of the progress of mathematics; instead it shows that art and science play off each other and that they are interlocked in a complex way. Long before the “discovery” of non-Euclidian mathematics by CarlFriedrich Gauß, Nicolaj Lobaschefskij, János Bolyai or Bernhard Riemann, art was already going in that direction by literally taking pictorial representation to the boundaries of Euclidian geometry, albeit without knowing what lay beyond or being able to imagine it. However, by constructing visible miracles it displayed the dispositive of perspective and opened up thinking about the conceivability of other spaces, even if they did not have any clear concept of what these might be. As the art of the fifteenth century foreshadowed developments which the mathematics of the seventeenth century caught up with, so the anamorphosis of the sixteenth and seventeenth century laid the foundations for a geometry whose fundamental and basic rules were only formulated and investigated one to two hundred years later. The fact that the nineteenth century showed little interest in anamorphosis, and that the twentieth century took up its strange trail to read the apotheosis of the fantastic from it, is essentially connected to what has been said above. In the work of Riemann and Felix Klein at the latest, the regime of Euclidian geometry as the only conceivable spatial system was broken, in the same way as technology overcame the per-
Dieter Mersch
36
ception of the subjective viewpoint, which had dominated the order of representation and its optics. Yet it appears that anamorphosis, looking far ahead and in an untimely way, dared to look at things “from the side,” which Roland Barthes in particular saw as its function and its special magic.37 It is as if one wanted, like Alice, to look behind the mirror in order to discover a completely different reality beyond the strict rule of the Euclidian system and rationally constructed space. It is nevertheless decisive that the possibility of such a discovery was already contained within the system itself from the beginning. Since the mathematical system and its technology clearly defined the strict realm of validity, it undermined its own invalidity. In so doing it established a norm and made its limits recognizable. At the limits a problem arose, however, a slanting angle as the breaking point towards the unreal or irrational, which could only appear unreal or irrational from the viewpoint of perspective because it did not seem describable within the laws of Euclidian space and its geometry. This is what the magic of anamorphosis means: its strange fascination consists in making the point at which the construction of the visible and its mathematics changed completely into its own decomposition visible. WORKS CITED Alberti, Leon Battista. On Painting. New York: Penguin, 1991. Baldus, R. and F. Löbell. Nichteuklidische Geometrie (= Sammlung Göschen, vol. 970/ 970a). Berlin: de Gruyter, 1964. Baltrušaitis, Jurgis. Introduction to Freed Leeman, Joost Elffers, and Mike Schuyt. Anamorphosen. Ein Spiel mit der Wahrnehmung, dem Schein und der Wirklichkeit. [Exhibit. cat.] Cologne: DuMont, 1975. Baltrušaitis, Jurgis. Anamorphoses ou perspectives curieuses. Paris: Perrin, 1955. [Expanded to: Anamorphoses ou Magie artificielle des effets merveilleux. Paris: Perrin, 2nd ed: 1969; new edition: Anamorphoses ou Thaumaturgus opticus. Paris: Flammarion, 1996.] Barthes, Roland. Criticism and Truth. Minneapolis: University of Minnesota Press, 1987. Becker, Oskar. Grundlagen der Mathematik in geschichtlicher Entwicklung. Frankfurt a.M.: Suhrkamp, 1975. Berger, John. Ways of Seeing. London: BBC/Penguin, 1972. Boehm, Gottfried. “Zwischen Auge und Hand. Bilder als Instrumente der Erkenntnis.” Mit dem Auge denken. Strategien der Sichtbarmachung in wissenschaftlichen und virtuellen Welten. Ed. Bettina Heintz and Jörg Huber. Zurich and New York: Springer, 2001. 43-54. 37
Roland Barthes. Criticism and Truth. Minneapolis: University of Minnesota Press, 1987. 80.
Representation and Distortion
37
Chastel, André, ed. Leonardo da Vinci. Sämtliche Gemälde und die Schriften zur Malerei. Munich: Schirmer/Mosel, 1990. Daston, Lorraine and Peter Galison. “Das Bild der Objektivität.” Ordnungen der Sichtbarkeit. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 29-99. Eser, Thomas. Schiefe Bilder. Die Zimmernsche Anamorphose und andere Augenspiele aus den Sammlungen des Germanischen Nationalmuseums. Nuremberg: Germanisches Nationalmuseum, 1998. Goodman, Nelson. Languages of Art. Indianapolis: Hacket, 1976. Haack, Wolfgang. Darstellende Geometrie (= Sammlung Göschen, vol. 144). Berlin: de Gruyter, 1969. Hauser, Arnold. Sozialgeschichte der Kunst und Literatur. Munich: C.H. Beck, 1969. Holländer, Hans. “Anamorphotische Perspektiven und Cartesianische Ornamente. Zu einigen Gemälden von Jean-François Nicéron.” Rezeption und Produktion zwischen 1570 und 1730. Festschrift für Günther Weydt. Ed. Wolfdietrich Rasch, Hans Geulen, and Klaus Haberkamm. Munich and Bern: Francke, 1972. 53-76. Holländer, Hans. “Mathematisch-mechanische Capriccios.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 347-54. Holländer, Hans. “Spielformen der Mathesis universalis.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 325-45. Kemp, Martin. Visualizations. The Nature Book of Art and Science. Berkeley: University of California Press, 2001. Krämer, Sybille. “Kalküle als Repräsentation. Zur Genese des operativen Symbolismus in der Neuzeit.” Räume des Wissens. Repräsentation, Codierung, Spur. Ed. Hans-Jörg Rheinberger, Michael Hagner, and Bettina Wahrig-Schmidt. Berlin: Akademie Verlag, 1997. 111-22. Krämer, Sybille. “Zentralperspektive, Kalkül, Virtuelle Realität: Sieben Thesen über die Weltbildimplikationen symbolischer Formen.” Medien-Welten, Wirklichkeiten. Ed. Gianni Vattimo and Wolfgang Welsch. Munich: Fink, 1998. 27-37. Lippe, Rudolf zur. Vom Leib zum Körper. Naturbeherrschung am Menschen in der Renaissance. Reinbek bei Hamburg: Rowohlt, 1988. Mansfeld, Jaap, ed. Die Vorsokratiker I. Stuttgart: Reclam, 1987. Mersch, Dieter. “Ästhetischer Augenblick und Gedächtnis der Kunst. Überlegungen zum Verhältnis von Zeit und Bild.” Die Medien der Künste. Beiträge zur Theorie des Darstellens. Ed. Dieter Mersch. Munich: Fink, 2003. 151-76. Meschkowski, Herbert, ed. Lust an der Erkenntnis. Moderne Mathematik. Munich and Zurich: Piper, 1991. Panofsky, Erwin. “Die Perspektive als ‘symbolische Form.’” Vorträge der Bibliothek Warburg 25 (1924), reprinted in H. Oberer and E. Verheyen, eds. Aufsätze und Grundfragen der Kunstwissenschaft. Berlin: Spiess, 1992. 258-330.
OLAF BREIDBACH
World Orders and Corporal Worlds: Robert Fludd’s Tableau of Knowing and its Representation World Orders World orders today are formed from knowledge, and as such they are determined by the subject. This idea is modern. In a world that was understood as creation, things were different. This world was determined in the absolute, which had nonetheless left its measure in the man created in its image. As such, this man was a reflection of the absolute and an image of the creation realized in himself. He became a model in which the whole of creation could then be seen. Man became a microcosm, a model of the whole, which was in turn determined by this model.1 The fact that this man was defined within a cosmic structure also structured his experience of the cosmos. The order of knowledge was not simply a reflection of what was conceived as possible, but something that appeared absolutely proportionate to this knowledge of the cosmos. As such, truth was not an unattainable greatness, sealed off from this man. Truth was to be found in the self, in the representation of a world developed within the self and – in one of the interpretations of the seventeenth century – that allowed one to find a tableau in this structure of the world system realized in it (fig. 1).2 In his form then, man maps out a model of the world, a microcosmos that lives within itself and that can encompass the macrocosmos in this life. Man can emerge within nature, nature can be illustrated in man. What does this model man offer in terms of a representation of knowledge? 1
2
Wilhelm Schmidt-Biggemann. Philosophia perennis. Historische Umrisse abendländischer Spiritualität in Antike, Mittelalter und Früher Neuzeit. Frankfurt a.M.: Suhrkamp, 1998. Cf. Harold J. Cook. “Physick and Natural History in Seventeenth-Century England.” Revolution and Continuity. Essays in the History and Philosophy of Early Modern Science. Ed. Peter Barker and Roger Ariew. Washington: Catholic University of America Press, 1991. 63-80.
World Orders and Corporal Worlds
39
Fig. 1: Robert Fludd. Utriusque Cosmi Maioris scilicet et MINORIS METAPHYSICA, PHYSICA ATQUE TECHNICA HISTORIA (Oppenheim, 1617).
This paper will sketch an order which allows man to be viewed a) through this form and b) through his understanding of this form. Models But what is a model? It is not simply an image, and it also cannot simply be explicated in language. The model has its own reality, in which constructions that are built upon it can be established. As a model, it provides the outlines of an idea, indeed an idea that is not completely translatable into a language of science, into a formula, or into a complex
40
Olaf Breidbach
sentence, and that cannot be explained by means of these formulations. As such, a medium for this idea is found in the model. The idea remains fixed to the forms of this medium, to what is technically possible and therefore also to the reality of this medium. At the same time, this way of viewing binds thought within the given form of viewing.3 It mediates a context within the image found with it. This model includes patterns of image and perception, patterns which are themselves a part of the idea, but that further influence its interactions with the model of reality found within it. This is perhaps a bit abstract, but this is precisely the difficulty that we find today in the area of models and simulation.4 The idea formed in the model determines itself. As such, its form and the formation explicated in it interweave in a way that is particularly meaningful for us. At the same time, thought does not simply offer a new aesthetic; the images found in this way are then simply shifted to a place before thought. The model fixes the idea to the visible. That these visual images then produce more than just popular illustrations of what science creates is the other side of such visualizability.5 These simple model images can also be used to view, evaluate, and interpret for and within the sciences. Images are not illustrations, they have their own reality. Athanasius Kircher, in his 1646 Ars magna lucis et umbrae, the very first media handbook, was clear about the fact that he was not describing a simple reflection of the world in the image’s worlds, but rather a reality in its own right.6 Here, incidentally, lies the decisive pictural turn, so important for our modern academic landscape, in which a culture of conversation became the book and with the book became illustration. But that is only incidental. The image not only presents an other that finds a new reality within itself, the image is first of all its own presentation. The image is within a frame in which it represents something.7 Furthermore, this does not change when the image moves and as such seems to come alive. 3 4 5
6
7
Cf. Olaf Breidbach. Das Anschauliche oder über die Anschauung von Welt. Vienna and New York: Springer, 2000. Cf. Rold Pfeifer and Christian Scheter. Understanding Intelligence. Cambridge and London: MIT Press, 1999. 59-78. Horst Bredekamp et al. “Bildwelten des Wissens.” Bilder in Prozessen (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1.1). Ed. idem and Gabriele Werner. Berlin: Akademie Verlag, 2003. 9-20. Athanasius Kircher. Ars magna lucis et umbrae in mundo, atque adeó universa natura, vires effectusque uti nova, ita varia novorum reconditiorumque speciminum exhibitione, ad varios mortalium usus, panduntur. Rome, 1646. Hans Belting. Bild-Anthropologie. Munich: Fink, 2001. 11-65.
World Orders and Corporal Worlds
41
Viewing Viewing an idea is therefore not a simple procedure. The patterns that result contain their own determinant. In 1908, Meyer’s lexicon defined the model simply as a “pattern, example: within construction a scaled down, prepared illustration, made of wood, clay, paper, plaster, cork, wax, etc., of an already constructed building or one yet to be constructed.” The model allows what is to be realized to be “assessed before its definitive creation.” This is especially important in the case of difficult construction materials.8 In 1976, this definition in Meyer will be expanded with the sentence “in academic usage a schematic, simplified, idealized representation.”9 Both of these citations refer to a story in which the viewing of the plaster model is only raised to this new, expanded level in modern science in the forms of its simulations. But is this story correct or not? The idea of such a model of thought is older than the natural sciences.10 Can we understand the complicated horologium sundials that were placed in the interiors of churches in the late Middle Ages as anything other than models of the cosmos?11 Isn’t even the clock – especially in the forms in which it was produced, for example, in the sixteenth century in Nuremberg – a model for measuring the world?12 And – true to Gottfried Wilhelm Leibniz’s view – the early adding machines are not simply arithmetical aids, rather they are a model of thought itself.13 Mathematical logic is transposed into them, in which – according to Leibniz – our thought is formalizable.14 These machines are not simply technologies that support human operations. They are in fact models in the above sense, explanations of 8 9 10 11
12 13 14
Meyers Großes Konversations-Lexikon. 6th ed. Leipzig and Vienna, 1908. Vol. 4, 12. Meyers Enzyklopädisches Lexikon. 9th ed. Mannheim, Vienna, and Zurich, 1976. Vol. 16, 364. Henning Genz. Gedankenexperimente. Weinheim: Wiley-VCH, 1999. Sebastian Münster. Fürmalung und künstlich beschreibung der Horologien. Basel, 1537; Gudrun Wolfschmidt. “Planeten, Kometen, Finsternisse – Peter Apian, Astronom und Instrumentenbauer.” Peter Apian. Astronomie, Kosmographie und Mathematik am Beginn der Neuzeit. Ed. Karl Röttel. 2nd edition. Buxheim: Polygon-Verlag, 1997. 93-106. Anthony John Turner. Of Time and Measurement. Studies in the History of Horology and Fine Technology. Aldershot: Variorum, 1993. Sybille Krämer. Berechenbare Vernunft. Kalkül und Rationalismus im 17. Jahrhundert. Berlin and New York: de Gruyter, 1991. Jörg Jochen Berns and Wolfgang Neuber, eds. Seelenmaschinen. Gattungstraditionen, Funktionen und Leistungsgrenzen der Mnemotechniken vom späten Mittelalter bis zum Beginn der Moderne. Vienna, Cologne, and Weimar: Böhlau, 2000.
42
Olaf Breidbach
ideas that in their function then also demonstrate a life of their own. By means of these machines, therefore, the Jesuits could demonstrate that not everything that functioned also had a soul and therefore was really alive. Even a preacher’s functions could be imitated by a machine.15 Jacques de Vaucanson’s mechanical duck, which enthralled the public in the 1740s, was also not simply a showpiece.16 This mechanical duck, which waddled across eighteenth century stages and fascinated audiences in Milan as late as the beginning of the nineteenth century, was a model of life. The fact that this duck, as well as moving, showed how chunks of food were masticated by its inner mechanics and, with the addition of fluid, were excreted in a completely new form, showed itself to be a model of the life process. A direct line can be traced from Vaucanson’s duck to Julien Offray de La Mettrie’s L’homme machine. In a work published in 1748, La Mettrie took the idea of the mechanical organization of human physiology to its logical conclusion in a theory that also took adding machines and their achievements seriously.17 His design for the comprehensive mechanization of human functions therefore – significantly – did not even stop at the spirit. The spirit could also be imagined as the function of a mechanism. It is interesting for us that these ideas were not conceived of as the result of physiological analyses, but rather as models, by means of which possible experiences of the organization of spiritual processes could then be ordered. Even here the model presents something that we had not yet understood, but it locates this something within its own reality. It is not unproblematic, however, that the model is already well established while our ideas are still feeling their way. The model therefore becomes a model of a kind of thinking that is trying to comprehend itself. Nor does it therefore stop at us: Accordingly, it would make perfect sense to write the history of logic as a model history of thinking about the self. Logic offers a model of how we might be able to think thought. In this way, thought gains an idea of itself.18 This idea, however, is limited to 15
16 17 18
Otto Mayr. “Automatenlegenden in der Spätrenaissance.” Technikgeschichte 41.1 (1974): 20-32; Horst Bredekamp. Antikensehnsucht und Maschinenglauben. Die Geschichte der Kunstkammer und die Zukunft der Kunstgeschichte. Berlin: Wagenbach, 2000. Peter Bexte and Werner Künzel. Maschinendenken/Denkmaschinen. An den Schaltstellen zweier Kulturen. Frankfurt a.M.: Insel Verlag, 1996. Julien Offray de La Mettrie. “L’homme machine.” Œuvres Philosophiques. 2nd edition. Berlin, 1774. Vol. 1, 275-356. Dieter Henrich. “Hegels Logik der Reflexion.” Die Wissenschaft der Logik und die Logik der Reflexion. Ed. idem. Bonn: Bouvier, 1978. 203-324; Rolf-Peter Horst-
World Orders and Corporal Worlds
43
what is possible within this model thinking. What thought knows about itself is very little. It contains schemes that become possible within such a schematism. It contains the patterns of a classification that – led by logic – point to a sequence of consequences, and it tries to bring every disconnected determinant into a series. With this idea, ways of organizing the cognitive come together, ideas that do formulate a sequence, but that then do not lead it out of the schema once it has been found. Even then, if I link this series and create a network out of them, this ‘statics’ of the system remains.19 The history of systems of managing knowledge illustrates how this only created new clothes for the old series of values.20 The ideas which lead these patterns of thought, ideas of the organization of knowledge, are therefore determined by old concepts that are only superficially spiced up in the image of a network and in the end do nothing other than maintain the reference structures of a library or the hierarchy of an encyclopedia.21 Networking The nets in which our knowledge is caught are here the nets of the old topoi. They are the thought patterns of Catholic cabalism. They are the ideas that allow for the organization of knowledge or – in the old way of speaking – the terms which convey this knowledge into a network that then allows the foundation of a library reference structure or an encyclopedic arrangement. Important figures marking this line of tradition are for example Leibniz, Kircher, or Johann Heinrich Alsted.22 Furthermore, it is a matter of the patterns in which theses ideas are caught. In Kircher, these are the presentations of networks that extract combination tables from themselves in the exact notation of a numerical sequence that unties the knot of the discursive in the tableau of the juxtaposition of the topos.23
19 20 21
22 23
mann. Ontologie und Relationen. Bradley, Russell und die Kontroverse über interne und externe Beziehungen. Königstein: Athenaeum, 1984. Cf. John Hertz, Anders Krogh, and Richard G. Palmer. Introduction to the Theory of Neural Computation. City Redwood: Perseus Books Group, 1991. Rüdiger Reinhardt. Wissen als Ressource. Theoretische Grundlagen, Methoden und Instrumente zur Erfassung von Wissen. Frankfurt a.M.: Peter Lang Verlag, 2002. Olaf Breidbach. “The Origin and Development of the Neurosciences.” Theory and Method in the Neurosciences. Ed. Peter K. Machamer, Rick Grush, and Peter McLaughlin. Pittsburgh: University of Pittsburgh Press, 2001. 7-29. Paolo Rossi. Clavis universalis. Arti della memoria e logica combinatoria da Lullo a Leibniz. Bologna: Il Mulino, 1983. Olaf Breidbach. “On the Representation of Knowledge in Athanasius Kircher.” Col-
44
Olaf Breidbach
The space that is encircled and patterned in this way is a first step towards such representational figures, that point in their abstraction to something higher. In his numerical rows, Kircher himself demonstrated that they refer to the infinite, but that they are still capable of being notated. If, in his volume on the Ars magna sciendi, a product of the possible combination theories as number takes up over a page in the book, the author in this way is proudly demonstrating the unmeasurability of what he has mastered in his schema.24 At the same time, this is also shown to be something that can be traced back to the imaginative space in human beings. As such, notation, and also the numeral, gain a dimension which reaches far beyond simple numerics. This numeral, which is described not as a mathematical quantity, but rather as the representation, seen almost as conjured up, of something illustrated in itself, is not only a unity on the scale of the possible, but the cosmos, broken back down into the imaginative space of the countable. The space of possibilities therefore once again almost acquires a human dimension in the numerical sequence. It is there for us to experience, to comprehend right on the page of a book. The cosmos of the topoi is understood as the inner space of our spirit.25 It acquires an order there that is represented as a spatial order. Our image of the world is formed in the inner stage that is constructed in this way. In 1550 Guilio Camillo built this inner space in a real theater.26 In this space, thought could view itself in the order that was possible for it. The world was staged in this theater; it could be experienced in its actual significance. Even Robert Fludd’s ‘Theatrum Mundi,’ whose representation Frances A. Yates wanted to use in the reconstruction of the Shakespearean Globe Theatre, is the image of an inner world of the spirit that itself becomes theater.27 The spirit becomes the stage of what is known to us. What we understand in this is nothing less than the world. And as such, it comes as no surprise that this I, which here becomes a world stage, is itself inscribed as an analog of the cosmos.28
24 25 26 27 28
lection, Laboratory, Theater. Scenes of Knowledge in the 17th Century. Ed. Helmar Schramm, Ludger Schwarte, and Jan Lazardzig. Berlin and New York: de Gruyter, 2005. 283-302. Athanasius Kircher. Ars magna sciendi . . . Amsterdam, 1669. Wilhelm Schmidt-Biggemann. Topica universalis. Eine Modellgeschichte humanistischer und barocker Wissenschaft. Hamburg: Meiner, 1983. Ludovico Domenichi. L’idea del theatro dell’ecc. M. Giulio Camillo. Florenz, 1550. Frances A. Yates. The Art of Memory. London: Routledge & Kegan Paul, 1966. Robert Fludd. Microcosmi historia. Tomus secundus de supernaturali, naturali et contranaturali microcosmi historia in tractatus tres distributa. Oppenheim, 1619.
World Orders and Corporal Worlds
45
This I corresponds to the outside: it is created as a depiction of God, and being a depiction of God, it is also a depiction of the world. It is the cosmos that God imagined. This finds its highest expression in the form in which God himself can be found in the world: in man. The God that bound himself in the form of man is God in this form, just as man was also created as the image of God.29 But what is this man in his similarity to God other than an image of the world, something that the world represents in itself? The old images, in which the world was revealed as bound in Christ, are the proof of what is represented in this form.30 The simple memorial culture of re-experiencing is expanded here to a living out of creation. Christ not only is the world, he bears it. And as a man, he bears it in himself. So man, at least within the possibility of Christ, becomes the cosmos. He becomes the measure and the foundation of the order that he understands as world order. His form, which corresponds to the form of Christ, is this measure of the cosmos.31 This measure is not outside the world, but is understood as emanation of the world itself. The world is understood as the measure of what is to be measured. Image Worlds The image worlds of Robert Fludd, which were published in the middle of the seventeenth century, show the consequences of such an approach to thought and the model. I am not engaging here with the theological program of these images.32 I want to look into the meaning that can possible be seen in them, rather than their genesis. For this reason, I take the image as representation, as a compression of the statements that, organized in the text, show a contour but not yet a profile of a world order. This will first be visible in the program of the images and in a quite particular sense will even be real in this observability. The man shown here is the Christ-man. The body of the world is to be found in him, which I find explicated in the construction of man. 29 30
31 32
Ibid. 71f.; cf. also the frontispiece of Robert Fludd. Anatomiae amphitheatrum effigie triplici, more et conditione varia designatum. Frankfurt, 1623. Hartmut Kugler. “Die Ebsdorfer Weltkarte. Ein europäisches Weltbild im deutschen Mittelalter.” Zeitschrift für deutsches Altertum und deutsche Literatur 116 (1987): 129. Fludd. Anatomiae amphitheatrum. Frontispiece. On this topic, cf. Frances A. Yates. The Occult Philosophy in the Elizabethan Age. London: Routledge & Kegan Paul, 1979.
Olaf Breidbach
46
Fig. 2: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
This body is therefore not abstract. It is the concrete body with its anatomy. This anatomy is put into a world order of numerically comprehensible analogies (fig. 2).33 The body of the world therefore can be inferred in an analysis of the cadaver’s configurations, accessible to the dissector, that can indeed hide an order within itself, but that has lost this order as such, that is to say, it can be dead and decayed. The man, however, at the same time also demonstrates in this frailness a mode of being of the cosmos that is precisely not absolute and eternal, but rather only existent in that it comes from God, and which is limited in its duration. Fludd himself considered this order to be a constantly returning time. He even situated it in the image of the information of the cosmos, of the assimilation of the world that can also be properly ‘shat’, as it is indeed nothing other than the excretion of God (fig. 3).34 33 34
Fludd. Anatomiae amphitheatrum. 111ff. Ibid. 27.
World Orders and Corporal Worlds
47
Fig. 3: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
On the other hand, Fludd is aware of the self-contained reaction of the organic, a reaction anchored in itself, which he believes that he comprehends in his ciphers of mechanical figures and in the image of a mechanism which is itself self-contained (fig. 4).35 All of this, however, I would now like to set aside and simply offer a few images that show how the world in Fludd is conceived from inside out, thereby turning man into a stage of the cosmos. This idea was not new, but astrological medicine lived by these analogies, just as much alchemy that sought to interpret its cures according to analogies in number and material relations of closed spaces of reaction.36 Fludd did not show man as a purely abstract form of possible analogies, but rather set him in relation to the idea of a God constructed as a 35
36
Robert Fludd. Utriusque cosmi maioris scilicet et minoris metaphysica, physica atque technica historia in duo volumina secundum cosmi differentiam divisa. Tomus primus. De macrocosmi historia in duos tractatus divisa. Oppenheim, 1617. 153ff. Charles John Samuel Thompson. The Lure and Romance of Alchemy. London: Harrap, 1932.
48
Olaf Breidbach
Fig. 4: Robert Fludd. Utriusque Cosmi Maioris scilicet et MINORIS METAPHYSICA, PHYSICA ATQUE TECHNICA HISTORIA (Oppenheim, 1617).
trinity. Man himself therefore – as the image of this God – becomes an existence with a threefold organization. He is seen as a being divided into head, trunk, and lower body. In this image of a being analogous to God, Fludd dismisses all the inessential particularities of human existence, such as the extremities inserted into this basic stratification. He draws the body as a space of reaction that stretches from the localized male reproductive organ of the lower body to the mental organ of the head. The body therefore becomes a tableau, a map, on which contours are drawn of the stratification of the human interior landscape, which can be related to the cosmic, which is analogous and equally organized by God. Both the cosmos and man prove to be existences joined according to a basic order. Therefore, this basic structure of the cosmos ‘man’, like the cosmos measured by the standard of man, is not to be found in a simple juxtaposition of parts, but is rather inserted into a hierarchy of relationships. The order to be traced here has both an above and a below. The above is determined from the part by which the world reaches into the heavenly, that is, the head, to which the lower levels relate, which then, however, in their reactions simultaneously explicate that this form, anchored in the heavenly, can also bring the heavenly into the world. The antithesis to this head and its ideas is the male reproductive organ, in that it allows for the creation of a new world, namely the creation of a new man (fig. 5).
World Orders and Corporal Worlds
49
Fig. 5: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
The being ‘man’, formed in this assimilation in terms of the idea of God accessible to the head, sets itself free through procreation in a world in order to arise anew from out of it (fig. 6). The head and the reproductive organ therefore give the poles a shape, which, in the balancing of the tension provided by this polarization, finds the way to itself, to its heart. If this inner tension is traced over both poles, it turns out that the pole is dissolved in its opposite. Fludd’s picture here is that of the pyramid that has its vertex in one of the poles and which opens itself up to the opposite pole. If I lay these two pyramids on top of one another, I get a middle, a fundamental something, a midpoint of the reactions in which the assimilation, leading to the head, and the dissimulation, connected with the opposite pole, overlap, and which allows assimilation to be transferred to dissimulation and vice versa. The central point in this area of overlapping pyramids is formed by the heart – or in the cosmic analogy– the sun, which warms the space of the possible (fig. 7).37 37
Fludd. Anatomiae amphitheatrum. 105ff.
Olaf Breidbach
50
Fig. 6: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
Human Topographies A human topography thus unfolds, which has its landmarks in the organs of the mind, of reproduction, and in the center of the circulation of assimilative and dissociative functions, in the heart. The three areas of the body defined by them are related to these organs.38 The trinity of human organization is as such an explication of the method of creation. All entities of the cosmos can be related to this tableau of basic human functions. As such, the body of man also becomes a stage of the world (fig. 8). These vertices of human topography show the map of cosmic functions and of the beings defined in this cosmos.39 The elements of the lower dissimulating levels, which assimilate themselves in the world bodies, are therefore to be intertwined in this functionally defined pattern of three, and as such, they are related to functions explicated in man. The 38 39
Ibid. 82f. Ibid. 259.
World Orders and Corporal Worlds
51
Fig. 7: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
basic elements of the organization of the world can then be arranged according to this schema of a Scientia pyramidalis, explicated in man, and can be brought into relation with each other.40 What is revealed is a panopticum of the world, which allows a connection in the ambiguous triple quality of the organizing structure between the analysis of the structures ‘physiognomia, chiromantia, and pyramidum scientia’ and the analysis of the functions ‘ars memoria, geomantia, and phophetia’ in the microcosmi historia, that is, in man.41 Man is then here the structure that a) allows the analytics of the structures to be created in themselves and at the same time b) – in the analysis of the functions of its analytics of this structure – allows it to be il40 41
Ibid. 80ff. Ibid. Tractatus primi, sectio secunda, 1.
52
Fig. 8: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
Olaf Breidbach
World Orders and Corporal Worlds
53
lustrated in a functional structure. Man is then also the figuration, on Earth, in the animality of the possible life that has grown in it, that breaks through this world structure in itself and thrusts forward through the world of the cosmos thus determined in itself into the absolute. That is how this is explained as a trinity. And so the triple-membered nature of the human being as a reflection of the forming of the absolute is demonstrated. What is shown in the organizational schema of the human being is a world body, which is revealed to be in man at the same time that it is also a reflection of the existence determined in itself, in which this cosmos was made possible in the first place. What is described is therefore an architecture of the human, which can be determined as the stage on which the structure of the world is modulated. In the memory, this world stage will be re-modulated. The ars memoria forms a theater in which the trinary organization of the human being can be staged within that being itself.42 By means of inner experience the structure of the world is comprehensible for man himself. He forms it according to his own architecture, which, determined as such in itself, has the human develop into the self-explicated measure of the cosmos of possible determinants. In this human being, this world can be represented in the multiplicity of its possible relationships, according to the organization of each determinant of the absolute in its triple membership. What is thus created in the topography of the body is the inner stage, on which the memorized outside world is created according to the triple quality of the function of memorizing as well. The experience of the memory grows in the classification of sensibility, imagination, and intelligibility.43 In the process, its sensibility is taken through the five senses. Accordingly, the inner stage of experience has five entrances and is evaluated from three positions.44 Thus, inner experience is a reflection of the organization of man that refers back to the cosmos. This is nothing other than the reflection and therefore also the realization of external harmony. Fludd developed – based on numerical magic – the schema of a comprehensive correspondence of events in the cosmos, which he then connected back to the organization of the human body and structured in terms of this body. 42
43 44
Cf. Wilhelm Schmidt-Biggemann “Robert Fludds Theatrum memoriae.” Ars memorativa. Zur kulturgeschichtlichen Bedeutung der Gedächtniskunst 1400-1750. Ed. Jörg Jochen Berns and Wolfgang Neuber. Tübingen: Niemeyer, 1993. 154-70. Fludd. Anatomiae amphitheatrum. 212ff. Ibid. Tractatus primi, sectio secunda, 55.
Olaf Breidbach
54
Fig. 9: Robert Fludd. Microcosmi historia (Oppenheim, 1619).
The point of departure is the divine number of the trias, whose explanation is contained in the symmetry of numbers and can then be decomposed into the Scientia pyramidalis on the body understood as a triple structure. Through this process, the multiplicity of the external can be understood and represented in its harmony. Man strives in his form to come out of the indeterminacy of the microcosmos – of the world where he has gained a foothold – into the light, the day of the microcosmos, which is to say God (fig. 9).45 As such, the microcosmos becomes a representation of the cosmic. Fludd formulates this in his representation De technica Microcosmi historia.46 45 46
Ibid. Tractatus primi, sectio prima, 275. Ibid. Tractatus primi, sectio secunda, 1.
World Orders and Corporal Worlds
55
And so it is no wonder that even the representation of the humanae scientia is organized in three books – and as such in the representation of the represented remains in the form in which the truth of the cosmos is to be found.47 In this context, to recognize is to enlighten. Sense is the prophecy acquired from the light of the microcosmos, from God, in which the pattern of the structuring of the cosmic is grasped, in which the rationality of this prophecy is explicated, and which then, in this explication, is realized as true.48 In the numerical mechanics of this classificatory calculation, a program of images of knowledge unfolds. In this then, the microcosm becomes a demonstration of the formation of the macrocosm by means of the Scientia Pyramidalis. The pyramid is therefore not a randomly chosen form, which only provides an image of the triple organization of each microcosmic reaction. This pyramid is much more the reality in which the details of the microscopic and macroscopic are first formed. The pyramid is the ray of the creating sun.49 It is itself the basic form in which creation can not only be represented, but rather that in which it can be initially realized at all. That this basic form of creation can then be experienced also, and particularly, in the shape of man and can be explicated in its structuring power, is hardly surprising any more. What is to be demonstrated then, is not simply illustrations of a theory of creation. Rather, it is the educational program that rests on a trinary structure in the form of the pyramid and explicated by this, a program in which the structuring of each formation – the information of the world – is determined as a size measured in human standards. Admittedly, this kind of representation is not accounted for in this way. An explanation in these terms – according to Fludd – also becomes obsolete, but the foundation for the construction of this illustrated world is seen in the reconstruction of the possibility of a pyramidic structure of the real. The explanation itself becomes real because it is shown to be possible in the image. The image constitutes the world. The image does not open a stage of possible classifications, it does not demonstrate a combinatorics of the artificial. The stage that can be understood in the image of man is itself the world. What is demonstrated is the world theater that is realized by analogizing microcosmos and macrocosmos according to the pattern of the Scientia pyramidalis. And so man becomes the measure of the world. The world becomes analyzable through him. In the arrangement of the elements of the world 47 48 49
Ibid. 3. Ibid. 179. Ibid. 248f.
Olaf Breidbach
56
in the bodies of men, the idea of a world made concrete in the shape of man and therefore made accessible to our analysis is achieved. Man becomes a model of the world. He becomes this, because God realized himself in this world in him, and so made the cosmos absolute in man. Naturalizations What is suggested in the analysis above is the basic moments of Fludd’s topography of the human. In this topography, the body of man became a stage of the cosmos. The structure of the cosmos – according to Fludd – was in this respect not only determined in man, it was – on the stage of memory – also explicitly realized in him. In the section on the body, Fludd creates a model that allows him to construct the world using the forms that he had discovered. For Fludd, then, the reality of the image became a starting point for extensive reflections. He offers a mode of construction that is explicated in the body of man. The image of the body becomes the measure by which the multiplicity of possible relationships in such a construction is positioned. The organism of man maps nature, seen on three levels, that is found presented in the “three cavities of the human body, so that in one of each of these spaces, one of the three spheres predominates over the others.”50 It is important to differentiate between the lower material sphere of the lower body, which is represented by the intestines, and the middle organic sphere of the trunk, in the center of which is the heart, and the upper psychic sphere of the head.51 These three spheres are therefore stretched between two poles, the upper pole of imagination, which corresponds to the brain, and the lower pole of reproduction, which corresponds to the sexual organ. The passages cited here are then not simply excerpts from Fludd. These lines come rather from the anthropology of Karl Friedrich Burdach in the middle of the nineteenth century.52 Burdach did not merely adopt Fludd’s basic premises. He follows, even in the details of his description, Fludd’s pattern of images, which we have indicated above. Burdach particularly adopted an understanding of the form of man as 50
51 52
Karl Friedrich Burdach. Beyträge zur näheren Kenntnis des Gehirns in Hinsicht auf Physiologie, Medicin und Chirurgie. First Part. Leipzig: Breitkopf und Härtel, 1806. 42. Ibid. 42ff. Olaf Breidbach. “Karl Friedrich Burdach.” Naturphilosophie nach Schelling. Ed. Thomas Bach and idem. Stuttgart-Bad Cannstatt: Frommann-Holzboog, 2005. 73-105.
World Orders and Corporal Worlds
57
trinary and as a hierarchically organized existence, which to this extent then also takes on the structure of the macrocosmos. Even more clearly than Lorenz Oken, who was working at nearly the same time, this ‘Romantic’ points to a European tradition of thought, with roots in the seventeenth century, whose detailed statements he used for the systematization of his own understanding of nature.53 The idea of structuring the world in this way according to the pattern of man had here become an independently existing model. This thought no longer needed to point to creation and therefore to some absolute situated outside nature. The pattern of nature was instead realized in itself, in the arrangement of the patterns of images created by it, and of the order formed within them. Oken names figures that now also continue to have an effect in arguments that are detached from Fludd’s way of understanding the cosmological.54 The de-theologized Fludd presented here is the Fludd of a theatrical reality, in which the acting found in the model becomes its own reality, and the world becomes a re-playing of this reality, which has been opened up on the stage of the human. Oken – after 1805 the champion of Friedrich Wilhelm Joseph Schelling55 in natural philosophy – goes further in the naturalization of such a cosmic numerology, set in human terms. He is also familiar with the schematics of the trinary. For him as well, man is the microcosmos in which the structure of the cosmos can be experienced. Man himself, however, is here understood as the highest realization of nature. In his structure, he does not point to anything higher, but rather to nature, realized in him in its highest form. It is then only logical that the other forms of nature are simply further elements of this nature, perfected in man and explicated in all possible specifications. Oken’s conclusion was also then to establish a systematics of animals as an applied anatomy of man.56 The forms of animals were nothing more than the organs of man made autonomous. Their nature is then ultimately only realizable in man. 53
54 55
56
On Oken cf. Olaf Breidbach and Michael T. Ghiselin. “Lorenz Oken’s ‘Naturphilosophie’ in Jena, Paris, and London.” History and Philosophy of the Life Sciences 24 (2002): 219-47. Michael T. Ghiselin. “Lorenz Oken.” Naturphilosophie nach Schelling. Ed. Thomas Bach and Olaf Breidbach. Stuttgart-Bad Cannstatt: Frommann-Holzboog, 2005. Thomas Bach. “‘Was ist das Thierreich anders als der anatomirte Mensch …?’ Oken in Göttingen (1805-1807).” Lorenz Oken (1779-1851). Ein politischer Naturphilosoph. Ed. Olaf Breidbach, Hans-Joachim Fliedner, and Klaus Ries. Weimar: Böhlau, 2001. 73-91. Olaf Breidbach. “Oken in der Wissenschaftsgeschichte des 19. Jahrhunderts.” Lorenz Oken (1779-1851). Ein politischer Naturphilosoph. Ed. idem, Hans-Joachim Fliedner, and Klaus Ries. Weimar: Böhlau, 2001. 15-32.
Olaf Breidbach
58
Oken also uses the form of the pyramid, for him determined in man, to measure the natural. The consequences of such a removal of a theologically determined stage in the reality of the human are familiar to us: The Scientia pyramidalis led in the nineteenth century to racial doctrines and to social Darwinism, which, by identifying the peak of the pyramid, called it the highest point and therefore that which encompassed everything else.57 It then becomes important to treat this peak correctly. In Oken, this identification of the highest possible organization, human and therefore also natural, was explicit. At the peak of his pyramid was the Prussian officer.58 The peak of human activity was then for Oken the art of war. With this, we are indeed – historically – far from Fludd himself, but we are still moving within the patterns of images sketched above. The topography, here attributed to Fludd, of a human who finds himself to be the measure of all things, continues to exert an influence. The concept of images is at any rate naturalized. Man, in Fludd still understood as coming from and heading towards God, becomes a final measure of the representation of the cosmos within an idea that determines nature itself as being included in it and does not provide for an understanding of nature as coming from God. Schemas According to Fludd, the inner representation of the world points to a higher being. If this is absent, the measure of the human thus conceived lacks a final anchor.59 Oken’s ideas, aimed at understanding nature in man, break down at any rate in Burdach, who positions this man only as a multiplicity of reactions within nature. Any idea of an absolute thus remains lost. The position of the measure itself remains open. After 1859, this idea of a measure determined in and of itself was converted into a theory of evolution. Nevertheless, the pattern of images of the inner world stage continues to exist. At least in the organization of knowledge and the concepts of memory, the old schema of an architecture secured in the human persists. This topos of the inner image therefore also became dominant in our own ideas about classification and order. Even 57 58 59
Emanuel Rádl. Geschichte der Biologischen Theorien in der Neuzeit. 2 vols. Leipzig, 1909. Lorenz Oken. Lehrbuch der Naturphilosophie. Zürich: Schultheß, 1843. 523. Cf. Olaf Breidbach. Deutungen. Zur philosophischen Dimension der internen Repräsentation. Weilerswist: Velbrück Wissenschaft, 2001.
World Orders and Corporal Worlds
59
modern expert systems are nothing more than such inner structures projected outwards.60 And even the experiments of twentieth century physiologists like John O’Keefe or Wilder Penfield, who tried to trace an inner landscape of experience in the brain, are connected as such to this old tradition.61
Translation: Daniel Hendrickson WORKS CITED Bach, Thomas. “‘Was ist das Thierreich anders als der anatomirte Mensch …?’ Oken in Göttingen (1805-1807).” Lorenz Oken (1779-1851). Ein politischer Naturphilosoph. Ed. Olaf Breidbach, Hans-Joachim Fliedner, and Klaus Ries. Weimar: Böhlau, 2001. 73-91. Belting, Hans. Bild-Anthropologie. Munich: Fink, 2001. Berns, Jörg Jochen and Wolfgang Neuber, eds. Seelenmaschinen. Gattungstraditionen, Funktionen und Leistungsgrenzen der Mnemotechniken vom späten Mittelalter bis zum Beginn der Moderne. Vienna, Cologne, and Weimar: Böhlau, 2000. Bexte, Peter and Werner Künzel. Maschinendenken/Denkmaschinen. An den Schaltstellen zweier Kulturen. Frankfurt a.M.: Insel Verlag, 1996. Bredekamp, Horst. Antikensehnsucht und Maschinenglauben. Die Geschichte der Kunstkammer und die Zukunft der Kunstgeschichte. Berlin: Wagenbach, 2000. Bredekamp, Horst et al. “Bildwelten des Wissens.” Bilder in Prozessen (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1.1). Ed. idem and Gabriele Werner. Berlin: Akademie Verlag, 2003. 9-20. Breidbach, Olaf. Die Materialisierung des Ichs. Zur Geschichte der Hirnforschung im 19. und 20. Jahrhundert. Frankfurt a.M.: Suhrkamp, 1997. Breidbach, Olaf. Das Anschauliche oder über die Anschauung von Welt. Vienna and New York: Springer, 2000. Breidbach, Olaf. Deutungen. Zur philosophischen Dimension der internen Repräsentation. Weilerswist: Velbrück Wissenschaft, 2001. Breidbach, Olaf. “Hirn und Bewußtsein. Überlegungen zu einer Geschichte der Neurowissenschaften.” Neurowissenschaften und Philosophie. Ed. Michael Pauen and Gerhard Roth. Munich: Fink, 2001. 11-58. Breidbach, Olaf. “Oken in der Wissenschaftsgeschichte des 19. Jahrhunderts.” Lorenz Oken (1779-1851). Ein politischer Naturphilosoph. Ed. idem, Hans-Joachim Fliedner, and Klaus Ries. Weimar: Böhlau, 2001. 15-32. Breidbach, Olaf. “The Origin and Development of the Neurosciences.” Theory and 60
61
Olaf Breidbach. “Hirn und Bewußtsein. Überlegungen zu einer Geschichte der Neurowissenschaften.” Neurowissenschaften und Philosophie. Ed. Michael Pauen and Gerhard Roth. Munich: Fink, 2001. 11-58. On this, cf. Olaf Breidbach. Die Materialisierung des Ichs. Zur Geschichte der Hirnforschung im 19. und 20. Jahrhundert. Frankfurt a.M.: Suhrkamp, 1997. 308ff.
60
Olaf Breidbach
Method in the Neurosciences. Ed. Peter K. Machamer, Rick Grush, and Peter McLaughlin. Pittsburgh: University of Pittsburgh Press, 2001. 7-29. Breidbach, Olaf and Michael T. Ghiselin. “Lorenz Oken’s ‘Naturphilosophie’ in Jena, Paris, and London.” History and Philosophy of the Life Sciences 24 (2002): 219-47. Breidbach, Olaf. “Karl Friedrich Burdach.” Naturphilosophie nach Schelling. Ed. Thomas Bach and idem. Stuttgart-Bad Cannstatt: Frommann-Holzboog, 2005. 73-105. Breidbach, Olaf. “On the Representation of Knowledge in Athanasius Kircher.” Collection, Laboratory, Theater. Scenes of Knowledge in the 17th Century. Ed. Helmar Schramm, Ludger Schwarte, and Jan Lazardzig. Berlin and New York: de Gruyter, 2005. 283-302. Burdach, Karl Friedrich. Beyträge zur näheren Kenntnis des Gehirns in Hinsicht auf Physiologie, Medicin und Chirurgie. First Part. Leipzig: Breitkopf und Härtel, 1806. Cook, Harold J. “Physick and Natural History in Seventeenth-Century England.” Revolution and Continuity. Essays in the History and Philosophy of Early Modern Science. Ed. Peter Barker and Roger Ariew. Washington: Catholic University of America Press, 1991. 63-80. Domenichi, Ludovico. L’idea del theatro dell’ecc. M. Giulio Camillo. Florenz, 1550. Fludd, Robert. Utriusque cosmi maioris scilicet et minoris metaphysica, physica atque technica historia in duo volumina secundum cosmi differentiam divisa. Tomus primus. De macrocosmi historia in duos tractatus divisa. Oppenheim, 1617. Fludd, Robert. Microcosmi historia. Tomus secundus de supernaturali, naturali et contranaturali microcosmi historia in tractatus tres distributa. Oppenheim, 1619. Fludd, Robert. Anatomiae amphitheatrum effigie triplici, more et conditione varia designatum. Frankfurt, 1623. Genz, Henning. Gedankenexperimente. Weinheim: Wiley-VCH, 1999. Ghiselin, Michael T. “Lorenz Oken.” Naturphilosophie nach Schelling. Ed. Thomas Bach and Olaf Breidbach. Stuttgart-Bad Cannstatt: Frommann-Holzboog, 2005. Hertz, John, Anders Krogh, and Richard G. Palmer. Introduction to the Theory of Neural Computation. City Redwood: Perseus Books Group, 1991. Henrich, Dieter. “Hegels Logik der Reflexion.” Die Wissenschaft der Logik und die Logik der Reflexion. Ed. idem. Bonn: Bouvier, 1978. 203-324. Horstmann, Rolf-Peter. Ontologie und Relationen. Bradley, Russell und die Kontroverse über interne und externe Beziehungen. Königstein: Athenaeum, 1984. Kircher, Athanasius. Ars magna lucis et umbrae in mundo, atque adeó universa natura, vires effectusque uti nova, ita varia novorum reconditiorumque specimi-num exhibitione, ad varios mortalium usus, panduntur. Rome, 1646. Kircher, Athanasius. Ars magna sciendi. In XII libros digesta, qua nova et universali methodo per artificiosum combinationum contextum de omni re proposita plurimis et propre infinitis rationibus disputari, omniumque summaria quaedam cognitio comparari potest. Amsterdam, 1669. Krämer, Sybille. Berechenbare Vernunft. Kalkül und Rationalismus im 17. Jahrhundert. Berlin and New York: de Gruyter, 1991. Kugler, Hartmut. “Die Ebsdorfer Weltkarte. Ein europäisches Weltbild im deutschen Mittelalter.” Zeitschrift für deutsches Altertum und deutsche Literatur 116 (1987): 1-29. La Mettrie, Julien Offray de. “L’homme machine.” Œuvres Philosophiques. 2nd ed. Berlin, 1774. Vol. 1, 275-356. Mayr, Otto. “Automatenlegenden in der Spätrenaissance.” Technikgeschichte 41.1 (1974): 20-32. Meyers Großes Konversations-Lexikon. 6th ed. Leipzig and Vienna, 1908. Meyers Enzyklopädisches Lexikon. 9th ed. Mannheim, Vienna, and Zurich, 1976.
World Orders and Corporal Worlds
61
Münster, Sebastian. Fürmalung und künstlich beschreibung der Horologien. Basel, 1537. Oken, Lorenz. Lehrbuch der Naturphilosophie. Zürich: Schultheß, 1843. Pfeifer, Rold and Christian Scheter. Understanding Intelligence. Cambridge and London: MIT Press, 1999. Rádl, Emanuel. Geschichte der Biologischen Theorien in der Neuzeit. 2 vols. Leipzig, 1909. Reinhardt, Rüdiger. Wissen als Ressource. Theoretische Grundlagen, Methoden und Instrumente zur Erfassung von Wissen. Frankfurt a.M.: Peter Lang Verlag, 2002. Rossi, Paolo. Clavis universalis. Arti della memoria e logica combinatoria da Lullo a Leibniz. Bologna: Il Mulino, 1983. Schmidt-Biggemann, Wilhelm. Topica universalis. Eine Modellgeschichte humanistischer und barocker Wissenschaft. Hamburg: Meiner, 1983. Schmidt-Biggemann, Wilhelm. “Robert Fludds Theatrum memoriae.” Ars memorativa. Zur kulturgeschichtlichen Bedeutung der Gedächtniskunst 1400-1750. Ed. Jörg Jochen Berns and Wolfgang Neuber. Tübingen: Niemeyer, 1993. 154-70. Schmidt-Biggemann, Wilhelm. Philosophia perennis. Historische Umrisse abendländischer Spiritualität in Antike, Mittelalter und Früher Neuzeit. Frankfurt a.M.: Suhrkamp, 1998. Thompson, Charles John Samuel. The Lure and Romance of Alchemy. London: Harrap, 1932. Turner, Anthony John. Of Time and Measurement. Studies in the History of Horology and Fine Technology. Aldershot: Variorum, 1993. Wolfschmidt, Gudrun. “Planeten, Kometen, Finsternisse – Peter Apian, Astronom und Instrumentenbauer.” Peter Apian. Astronomie, Kosmographie und Mathematik am Beginn der Neuzeit. Ed. Karl Röttel. 2nd edition. Buxheim: Polygon-Verlag, 1997. 93-106. Yates, Frances A. The Art of Memory. London: Routledge & Kegan Paul, 1966. Yates, Frances A. The Occult Philosophy in the Elizabethan Age. London: Routledge & Kegan Paul, 1979.
FLORIAN NELLE
Telescope, Theater, and the Instrumental Revelation of New Worlds Instruments have a double function in the seventeenth century. They are a source of wonder and at the same time are intended to aid orientation in the new reality that they reveal. This applies to scientific instruments such as the telescope and microscope, which reveal a new nature. It applies to poetic instruments like the mannerist metaphor, which is meant to open up a new form of intellectual experience. However, it applies especially to instruments of political culture, which reveal new forms of society. In particular, this includes the theater of machines, which not only visualizes the transformation of literary myths into the world of the absolutist state, but also performs it on stage. Thus, an instrument for the manifest staging of a new world is established beyond the mere visuality of writings and pictures. This accessible utopia can become the scene for stagings of faith, power, and science. The development of the theater of machines as an institution and place of power happens at the time of the world’s instrumental development by means of the use of the telescope, the microscope, and later the pneumatic pump. The importance of this new form of theater for cultural politics is practically equivalent to the significance of the new instruments for the philosophy of science. It is evident that there is a connection between the instruments of science and those of politics. The telescope, the microscope, and the pneumatic pump as well as the technically advanced theater of the time are regarded as instruments of revelation. According to Robert Hooke in his 1668 Micrographia, the telescope and the microscope expanded the “world of senses”: “Hence there is a new visible World discovered to the understanding.”1 How1
Robert Hooke. “Preface.” Micrographia or some Physiological Descriptions on Minute Bodies made by Magnifying Glasses. New York: Dover, 1961 [facsimile of the edition London, 1665]. For the radical epistemological break caused by the invention of the telescope, cf. Hans Blumenberg. Der Prozeß der theoretischen Neugierde. Frankfurt a.M., 1973. 170.
Telescope, Theater, and the Instrumental Revelation of New Worlds
63
ever, similar statements can be found about the theater of machines, as in Abbé d’Aubignac’s 1657 Pratique de Theatre: “[The theater of machines] shows us a new sky, a new earth, and many other wonders which we believe we see while knowing we are deceived.”2 In fact, discovering new worlds was a major topic of discussion in the seventeenth century. As early as 1644, René Descartes invited readers of Le Monde to “wander beyond this world to view another, wholly new, world, which I call forth in imaginary spaces.”3 In 1668, John Dryden claimed that in the course of the previous decade he “almost saw a new nature uncovered.”4 The revelation of new worlds was a contemporary trend. This is not only connected to European expansion, which had led to the discovery, conquest, and destruction of America over a hundred years earlier. It is also related to the idea of man’s resemblance to God. “I saw a new heaven and a new earth,” are the words of the prophet who looks upon the heavenly Jerusalem in the Revelation of John. This biblical background is always present whenever new worlds are mentioned. In 1607, Federico Zuccari explicitly formulates the above idea. He writes that man is to a certain extent a “second God.” Due to this “divine spark,” he is provided with the ability to create a “new world” for himself. As he “imitates God and emulates nature,” he is able to “produce an infinite number of artificial things which are similar to nature. With the aid of painting and sculpture, he can produce new paradises on earth.”5 Although Zuccari is here referring to the fine arts, he explicitly connects this ability to create new worlds with science and technology. However, he is not as concrete as Francis Bacon, who places laboratories 2
3
4 5
“Il est certain que les ornemens de la Scéne font les plus sensibles charmes de cette ingenieuse Magie . . . qui nous met en vuë un nouveau Ciel, une nouvelle Terre, & une infinité de merveilles que nous croions avoir présentes, dans le temps même que nous sommes bien assûrez qu’on nous trompe.” François Hédelin Abbé d’Aubignac. La pratique du théâtre. Munich: Fink, 1971. 319. René Descartes. The World and Other Writings. Ed. Stephen Gaukroger. Cambridge: Cambridge University Press, 1998. 21; cf. René Descartes. “Traité de la lumière.” Œuvres de Descartes. 13 vols. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1969. Vol. 11, 31. John Dryden. The Works. 20 vols. Ed. Edward Niles Hooker. Berkeley, Los Angeles, and London: University of California Press, 1971. Vol. 17, 15. “E fosse quasi un secondo Dio, volle anco dargli facoltà di formare in sé medesimo un dissegno interno intelletivo, accioché col mezzo di questo conoscesse tutte le creature e formasse in sè stesso un novo mondo . . . et inoltré accioché con questo dissegno, quasi imitando Dio et emulando la natura, potesse produrre infinite cose artificiali simili alle naturali, e col mezo della pittura e della scoltura farci vedere in terra novi paradisi.” Federico Zuccari. “Idea de’ pittori, scultori et architetti.” Scritti d’Arte del Cinquecento. Ed. Paola Barocchi. Milan: Ricciardi, 1973. Vol. 2, 2069.
64
Florian Nelle
and workshops at the center of his 1627 New Atlantis. Their purpose is to explore and reproduce every field of human and divine creation, and they are especially able to reproduce meteorological phenomena.6 Literally as well as figuratively, the program here is to create heaven on earth. Descartes takes up this idea in 1637. He invents a device to create rainbows artificially and uses it to make crosses and other “signs” appear in the sky. Naturally, they “arouse great astonishment among those who are unaware of the cause.”7 Thus, it may not be by chance that Descartes and Bernard Le Bovier de Fontenelle each take six days to create the new world of science. In Bacon’s New Atlantis “Solomon’s House” is called the “College of the Six Days Works.” For Descartes, it takes six meditations to make the world descend into nothing and be founded anew. Fontenelle needs six conversations to explain the Cartesian teaching on the structure of space properly and comprehensibly.8 Worlds are created in six days – with the aid of instruments of all kinds. Therefore, it is not surprising that for Hooke the refining and improvement of “the Sense, the Memory, and Reason” by means of instruments have a biblical background. It is nothing but a means of recovering the knowledge which established man's command over nature and which was lost with the Fall of man and the expulsion from paradise.9 As is indicated here, the purpose of instruments in the seventeenth century is more than finding facts and truths. Yet they certainly do fulfill this purpose. They are inquisitorial instruments, as they, in Hooke’s words, force nature “to confess, either directly or indirectly, the Truth of what we inquire.”10 Furthermore, they nurture hopes for the profitabil6 7
8
9
10
Francis Bacon. “New Atlantis.” The Major Works. Ed. Brian Vickers. Oxford: Oxford University Press, 1996. 480. “Et cecy me fait souvenir d’une invention pour faire paroistre des signes dans le ciel, qui pourroient causer grande admiration a ceux qui en ignoreroient les raisons.” Descartes. Œuvres de Descartes. Vol. 6, 343. Cf. René Descartes. “Meditations on First Philosophy.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdach. Cambridge: Cambridge University Press, 1984. Vol. 2, 1-62, and Bernard le Bovier de Fontenelle. Conversations on the Plurality of Worlds. Trans. H.A. Hargreaves. Berkeley and Los Angeles: University of California Press, 1990. (Thank you to Jan Lazardzig for pointing out the parallel.) “The only way which now remains for us to recover some degree of those former perfections, seems to be, by rectifying the operations of the Sense, the Memory, and Reason since upon the evidence, the strength, the integrity, and the right correspondence of all these . . . our command over things is to be establisht.” Hooke. Micrographia. N. p. “Such Experiments therfore, wherein Nature is as ‘t were put to Shifts and forc’d to confess, either directly or indirectly the Truth of what we inquire, are the best if
Telescope, Theater, and the Instrumental Revelation of New Worlds
65
ity of imperialism and colonialism. Hooke proudly emphasizes the way the Royal Society is financially supported by pragmatic businessmen. In this context, it should be mentioned that Bacon was also concerned with contemporary colonial projects. In 1597, he wrote the short memorandum Of Plantations, and in 1609 contributed to an expedition which ended up stranded in the Bermudas. (In his turn, William Shakespeare was inspired to take up the issue of colonization in The Tempest.) Finding a method to determine longitude, one of the main tasks of the Royal Society, was considered essential for maintaining England’s naval supremacy. As an anonymous poet states in a 1663 poem on the Royal Society: The Colledge will the whole world measure, Which most impossible conclude, And Navigators make a pleasure By finding out the longitude.11
For Robert Boyle, one of the more important purposes of the new sciences was colonizing the new world. Obviously, instruments in the seventeenth century have to be seen in this context of an inquisitorial quest for the truth, which was intended to benefit imperial politics as well as colonial trade. In addition, instruments are devices for the production of wonders. Visually perceptible phenomena become aesthetic phenomena. They are transformed into a drama that can be enjoyed as a work of art, stimulating not only reason but, as Hooke emphasizes, the senses as well. However, the underlying idea of aesthetics found here is obviously anachronistic. In a contemporary context, this phenomenon can most closely be described in terms of the meraviglia teachings of contemporary poetry.12 Just as the poet has to amaze the reader with surprising effects and unusual views, a look through the microscope reveals surprising and strange views of familiar items. Both Galileo Galilei’s telescopic views of the moon’s surface as well as Hooke’s microscopic examination of a needle tip reveal a remarkable fascination with the materiality of things. The discovery of the “roughness” of the moon’s surface revises the Aristotelian view of its
11 12
they could be met with.” Robert Hooke. The Posthumous Works. New York and London: Johnson, 1969 [facsimile of the edition London, 1705]. 34. Anonymous. “The Ballad of Gresham College.” Ed. Dorothy Stimson. Isis, 18 (1932): 103-17. Cf. Florian Nelle. “Im Rausch der Dinge. Poetik des Experiments im 17. Jahrhundert.” Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst. Ed. Helmar Schramm et al. Berlin: Dahlem University Press, 2003. 140-68, and Florian Nelle “Descartes und der Regenbogen im Wasserglas – Von der beobachteten zur inszenierten Natur.” Theatralität und die Krisen der Repräsentation. Ed. Erika Fischer-Lichte. Stuttgart and Weimar: Metzler Verlag, 2001. 392-410.
66
Florian Nelle
smoothness. For Galileo, in his Dialogue on the two world systems, the discovery of the roughness of the moon’s surface is the starting point for a description of the aesthetic features of the most varied surface structures. This contemplation and characterization of different surface qualities eventually culminates in a comparison of “sculptured velvet” and loosely wooded “mountain ridges.” In this way, Galileo tries to prove that the moon has mountains but no vegetation similar to Earth’s.13 Using this and other comparisons, Galileo’s interlocutors practice a poetic view of nature that is induced by instruments and irretrievably alters the way reality is viewed. This instrumental alienation effect becomes, in Hooke’s and Boyle’s work, a method of experimental observation with the purpose of producing surprising insights. They challenge their readers to approach everyday objects and things from the perspective of a traveler in a foreign country, to examine familiar items as if they were strange and unknown. Thus to a certain extent they transform the process of instrumental alienation into a method for scientific observation. Once this mode of observation is practiced, it can be used without the instrument. The human eye itself becomes an optical instrument and, as Boyle notes, one’s “alertness” becomes the “looking glass.”14 Thus man is able to gain essential insights from a plethora of seemingly meaningless events. Boyle illustrates the use of this method by an allegory that itself is the result of one such meditation. Just as a proficient craftsman is able to produce mirrors, binoculars, and burning glasses from ashes and sand, a “practiced observer [can put] these little fragments, or parcels of time [in such an order] as to afford us both looking-glasses to dress our souls by, and perspectives to discover heavenly wonders.”15 Thus, everyday things can become the stage for instrumentally sharpened curiosity. It is definitely possible to call this the poetics of the instrument, especially as this experimental art of observation is so strikingly close to the mannerist program. At the same time, the telescope and other optical glasses become the central emblem of mannerist poetics. In Emanuelo Tesauro’s work, Galileo’s discovery of sunspots thus becomes an allegory for a surprising twist, thanks to the clever poet’s acutezza. Like a telescope, the metaphor’s role in mannerist poetry is to connect remote points with each other. By bringing together seemingly 13 14 15
Galileo Galilei. Dialogue Concerning the Two Chief World Systems. Trans. Stillman Drake. Berkeley: University of California Press, 1953. 112. Robert Boyle. The Works. Ed. Thomas Birch. Hildesheim: Ulms, 1966. Vol. 2, 343 [facsimile of the edition London, 1772]. Ibid. 337.
Telescope, Theater, and the Instrumental Revelation of New Worlds
67
disparate points it creates surprising insights. Thus the instrument becomes an essential device for the production of wonders. It presents nature as an astonishing stage for a mysterious world, which – and this is actually the point – has to be penetrated by means of instruments. Insights created by instruments were far different from the true appearances of things, for several reasons.16 In the seventeenth century, every instrument was unique, and the crafting of the individual instruments could lead to different results. In the case of the microscope, for example, lighting was essential. Hooke himself cites an example where the surface of a fly’s eye seems different every time the light conditions under the microscope change. In addition, culturally determined interpretation also influenced the way things were perceived. Antoni van Leeuwenhoeks, for example, believed that he had discovered that the shape of human semen prefigures the human form. Yet it is crucial to know that, ultimately, instruments are not limited to revealing a new world. Hooke writes that “every considerable improvement” of instruments shifts the visible horizon and offers “new Worlds and TerraIncognita’s to our view.”17 The revelation moves beyond a single experimental act and becomes an infinite process. Far from providing certainty, this process presents a vast chaos of new worlds, in which instruments become an indispensable aid to orientation. Instruments not only dazzle by staging dramas of nature, they also become essential tools for solving the world’s riddles. Optic glasses, looking glasses, telescopes and microscopes become symbols of the fallibility of the senses. Paradoxically, this fallibility leads to a subtle compulsion to examine nature with instruments. Thus the revelation of new worlds by means of instruments presents a drama, which also puts the spotlight on the instruments themselves. Fascination moves from the object of knowledge, discovered using the instrument, to the instrument itself and thus to the person who created the instrument or knows how to use it. This is shown by Hooke’s decision to equip sense, memory, and reason with instruments. In passing, it should be mentioned that Hooke not only means sight when he mentions sense but hearing, feeling, taste, and smell as well. He thus anticipates the idea of flavor enhancers.18 In this vision, instruments are no longer merely tools for per16
17 18
Cf. Hartmut Böhme. “The Metaphysics of Phenomena. Telescope and Microscope in the Works of Goethe, Leeuwenhoek, and Hooke.” Collection, Laboratory, Theater. Scenes of Knowledge in 17th Century. Ed. Helmar Schramm, Ludger Schwarte, and Jan Lazardzig. Berlin: de Gruyter, 2005. 355-93. Hooke. Micrographia. N. p. Ibid.
Florian Nelle
68
ception. They change the organs of perception, just as the view through the telescope and microscope can lead to an instrumental view. A view that is fixed on the detail of reality provided by the instrument is replaced by a view fixed instead on the instrument and the person indissolubly merged with it. Thus, man is able to create an artificial paradise only as he becomes an artificial man with the aid of instruments. This becomes particularly impressive when the instrumental creation of a poetic view of reality is transferred to art. For example, painters in the late eighteenth century made use of “Claude Lorrain glasses.” In these tinted mirrors and glasses the landscape seems to be coated with the sweet veil of melancholic denial as in a soft-focus lens. The use of these glasses made Lorrain’s landscape paintings especially attractive.19 Bacon considered colored glasses and distorting mirrors the embodiment of the way human perception can be subject to deception. Now they had become a means of standardizing the examination of nature and of training one’s aesthetic eye. According to William Gilpin, having a trained eye enables man to grasp nature’s picturesque aspect and do without aids. “Nature gave us an apparatus that is more apt to see things in a picturesque light than any machine an optician could ever devise.”20 Instruments create aesthetic effects of wonder and reveal new worlds. They offer orientation in the chaos they cause, and yet they mostly refer to themselves. All of this is true for the theater as well, which at the turn of the seventeenth century becomes the prime instrument of cultural politics. Scientific instruments promise to create artificial paradises. Yet it is the theater of machines that constructs these new and better worlds in an exemplary way. Theater as an Instrument of Politics While the new scientific instruments in the seventeenth century opened up new spaces of visibility, theater became to the general model of cultural experience, becoming a universal instrument for explaining the world. From the courtier’s art of pretence21 to dramas of nature and his19
20 21
“The only picturesque glasses are those, which the artists call Claude Lorrain glasses. They’re combined of two or three different colours; and if the hues are well sorted, they give the objects of nature a soft, mellow tinge, like the colouring of that master.” William Gilpin. Observations on the Highlands of Scotland. Richmond: Richmond Publishing Co., 1973. 124 [facsimile of the edition London, 1789]. Ibid. Cf. August Buck. “Die Kunst der Verstellung im Zeitalter des Barocks.” Festschrift
Telescope, Theater, and the Instrumental Revelation of New Worlds
69
tories of the world and salvation, theater offered a platform for articulating and reflecting on individual, social, scientific, and eschatological questions.22 Manifested as the theater of machines, it entirely entered the service of aristocratic representation. The uniqueness of the representative function of the theater of machines was not only to be found in the myths it performed. (Inevitably, they were a variation on the golden age and the harmony of the spheres. Set pieces of antique myths were processed to become allegories of the virtuous reign of the prince who was to be celebrated.23) The technical and logistical apparatus necessary for the performance of the spectacle was at least as important. The theater of machines not only presented new worlds on stage, it also epitomized the ability to create them. Being the most complex medium of its time, it did indeed represent a microcosm of the world it reflected. It brought together visual and rhetorical art with lighting design and pyrotechnics. The craft of carpenters was connected with the mechanical designs and ingenious machines of engineers who used their knowledge of siege techniques, hydraulics, shipbuilding, and architecture. In other words, the theater of machines incorporated the latest developments in technology and provided a platform for carrying out experiments and innovations. From this point of view there is a very concrete background for the discussion of theatrum mundum. When it unfolds in all its magnificence, the theatrical machine of the seventeenth century summarizes the essential artistic and technical skills of its time. The stage engineer himself united a variety of skills. Bernardo Buontalenti, for example, was an architect, a builder of fortresses, and an inventor of hydraulic machines. Joseph Furttenbach worked as a theater engineer and was a master builder of fortresses. Francesco Guitti, who worked in Ferrara, was not only a theater architect, stage engineer, scenographer, and inventor of military and hydraulic machines but also a diplomat, poet, and spy.24 For a performance in 1644, Gianlorenzo Bernini “painted the scenes, cut the statues, invented the engines, com-
22
23 24
der Wissenschaftlichen Gesellschaft an der Johann Wolfgang Goethe-Universität Frankfurt a.M. Wiesbaden: Steiner, 1981. 85-103. Cf. Helmar Schramm. Karneval des Denkens. Theatralität im Spiegel philosophischer Texte des 16. und 17. Jahrhunderts. Berlin: Akademie Verlag, 1996. Concerning role metaphors cf. Ralf Konersmann. Der Schleier des Timanthes. Perspektiven der historischen Semantik. Frankfurt a.M.: Fischer, 1994. 84. Cf. Roy Strong. The Renaissance Garden in England. London: Thames and Hudson, 1998. Giuseppe Adami. “L’ingegnere-scenografo e l’ingegnere-venturiero.” Barocke Inszenierung. Ed. Rudolf Preimesberger and Joseph Imorde. Emsdetten and Zurich: Edition Imorde, 1999. 160.
70
Florian Nelle
posed the Musique writ the comedy & built the theater all by himself.”25 Such a stage engineer embodies Descartes’s idea of engineers, who “lay out [a new city] as they fancy on level ground” and thus avoid the mistakes of works “composed of several parts and produced by various different craftsmen.”26 He is closest to the ideal of the uomo universale, an ideal that was influenced by Vitruvius’ image of the ideal architect. This ideal architect was supposed to be a learned scholar and an expert in drawing and geometry. He should also be familiar with history, mythology, and philosophy, and should understand music and medicine, jurisprudence and astrology.27 As an archetypical creator of new worlds, the stage engineer was also regarded as a “quimérico ingeniero” or “grand sorcier,” such as Cosimo Lotti in Madrid or Giacomo Torelli in Paris.28 As technology as the embodiment of man’s creative power became immensely fascinating, a look behind the scenes proved more impressive than the drama on stage. In Paris, Torelli introduced a system using rope winches and counterweights that made the changing of scenery considerably easier. “This trick of transformation was wonderful because a single lad of fifteen years put it in motion . . . And really, without this ingenious trick it would have been hard to believe that so many flats move in an instant to change the scene.”29 The impression made by this spectacle can hardly be overestimated. It inspired a new dramatic genre. Lope de Vega, Pedro de Caldéron de la Barca, and Ben Johnson regarded their plays as the victims of spectacular effects. Pierre Corneille’s 1650 play Andromède, on the other hand, was especially written for Torelli’s newly invented effects. Corneille declared that the machines were a central part of the tragedy’s plot. As he explained in the preface: “Every act [has] its particular decoration and at least one flying ma25 26
27 28 29
John Evelyn. The Diary. 6 vols. Oxford: Clarendon Press, 1955. Vol. 2, 261. René Descartes. “Discourse on the Method.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdach. Cambridge: Cambridge University Press, 1984. Vol. 1, 116. Vitruvius. The Ten Books of Architecture. Trans. Ingrid D. Rowland. Cambridge: Cambridge University Press, 1999. Cf. N.D. Shergold. A History of the Spanish Stage from Medieval Times until the End of the Seventeenth Century. Oxford: Clarendon Press, 1967. 275 “Mirabile era l’artificio di questa mutatione, poiché un solo giovanetto di quindici anni le daba il moto . . . e in vero, ché fuori di questo ingegnoso artificio l’intelleto non s’adegua alla credenza, che tanti tellari possano tutti in un punto, in un baleno, anzi in un atomo accomodarsi a’ suoi luoghi per variare una Scena.” Quoted in Raimondo Guarino. La Tragedia e le macchine. Andromède di Corneille e Torelli. Rome: Bulzoni, 1982. 10.
Telescope, Theater, and the Instrumental Revelation of New Worlds
71
chine . . . In this tragedy, the machines are not detached decoration. They tie the knot and undo it, and they are so important that not a single one can be removed without causing the whole edifice to collapse.”30 Thomas Corneille even put machines in the titles of his plays. One of his pieces was published as Circè, tragédie, ornée de machines, de changements de théatre, & de musique.31 Thus the tragédie à machines itself became a category of drama. Even the presentation of what was most sacred could come second to the fascination for this perfectly organized theatrical machine. This applies to religious stagings of the Eucharist like the Quarantore machines of the counter-reformation. It also applies to the 1640 theatrum sacrum, where Niccolò Menghini “showed heaven on Earth to the people.”32 The perfectly organized bustling activity behind the scenes that were set up for the staging of the sacrament proved just as astonishing as the presentation of the most sacred itself. One saw scaffoldings and porches, which were built with extraordinary skill and symmetry, and skillfully laid boards and staircases. More than forty people were stationed in the finest order in the interior of this great machine. They lighted it and kept watch, so that no accident would happen, and not one of them could be seen by the audience. Many persons of high standing, who left this apparatus very pleased and astonished, can testify to this.”33
The motif of looking behind the scenes at the machinery of an artificial world quickly entered the Theatrum Mundi. It became one of the cen30
31 32
33
“Chaque acte . . . a sa décoration particulière, et du moins une machine volante . . . des machines, qui ne sont pas dans cette tragédie comme des agréments détachés, elles en font le noeud, et le dénouément, et y sont si nécessaires, que vous n’en sauriez retrancher aucune, que vous ne fassiez tomber tout l’édifice.” Pierre Corneille. Andromède. Tragédie. Ed. Cristian Delmas. Paris: Didier, 1974. 11. Cf. S. Wilma Holsboer. L’histoire de la mise en scène dans le théâtre français de 1600 à 1657. Paris: Droz, 1933. 151. Thomas Corneille. Circè, tragédie, ornée de machines, de changemens de théâtre, & de musique. Paris, 1690. “Niccolò Menghini Romano . . . hà superato se stesso in far veder à gli huomini quasi vn Cielo in terra.” Antonio Gerardi. “Relatione del Solenne Apparato.” Quoted in Joseph Imorde. Präsenz und Repräsentanz oder: die Kunst, den Leib Christi auszustellen. Das vierzigstündige Gebet von den Anfängen bis in das Pontifikat Innocenz X. Emsdetten and Berlin: Edition Imorde, 1997. 152. “Es se fusse stato possibile far veder’ à ciascuno l’interno, non minore meraviglia gli aveva recato, che l’esterno, poiche vedeuansi con architettura, e simmetria straodinaria diversi palchi, salti, risalti, distanze di tavole ben collocate, e scale, per le quali con bell’ordine, e sicurezza nell’immensità della vasta machina più di quaranta persone stavano distribuite illuminando, e facendo la sentinella al suo posto, acciò non seguisse danno alcuno, e niuno di questi fù mai veduto dagli spettatori: e di questo ch’io dico sono testimonij molti Personaggi grandi, che si sono compiaciuti di salire per tutto con loro gran gusto e stupore.” Ibid. 150.
72
Florian Nelle
tral elements in the theatrum’s model for explaining the world. Fontenelle used it to justify empirical philosophy. He compared the world to a play at the opera. While the audience is following the effects of stage machinery in astonishment, empirical philosophy looks at the winches, ropes, and weights that set the spectacle in motion. Fontenelle uses the winches and weights to visualize the hidden mechanism of the drama of nature. On the other hand, for Jean de la Bruyère, they signify the mechanism of self-violation. According to him, men have to make use of this mechanism on their way to success, whenever they present themselves in public. Looking into their souls reveals this mechanism. If you go behind the scenes at a theatre and count the weights, the wheels, the ropes which make the machines work and the actors fly, if you consider how many people are involved in making the things move, and what strength of arms, what muscular exertion they display, you will say: “Are these the principles and the secret springs of that splendid spectacle that seems so natural, as though it were alive and moved of its own accord?” You will protest: “What strenuous and violent efforts!”34
However, these are later variations on the motif. Yet when the view behind the scenes appeared for the first time, it undoubtedly pointed to the power of the monarch who funded the spectacle for his own glorification. The festivities on the occasion of the wedding of the Grand Duke Ferdinand of Tuscany in 1589 exemplify this well,35 especially Buontalenti’s stagings of the intermezzi from the opera La Pellegrina. Incidentally, they set new technical standards and created significant and fundamental elements of the baroque theater of machines.36 The subject that 34
35
36
“Si vous allez derrière un théâtre, et si vous nombrez les poids, les roues, les cordages, qui font les vols et les machines; si vous considérez combien de gens entrent dans l'exécution de ces mouvements, quelle force de bras, et quelle extension de nerfs ils y emploient, vous direz: ‘Sont-ce là les principes et les ressorts de ce spectacle si beau, si naturel, qui paraît animé et agir de soi-même?’ Vous vous récrierez: ‘Quels efforts! quelle violence!’” Jean de la Bruyère. Les caractères. Paris, 1976. “Des bien de fortune,” § 25. This celebration has the advantage of being one of the most documented and wellresearched of its time. Cf. James M. Saslow. The Medici Wedding of 1589. Florentine Festival as Theatrum Mundi. New Haven and London: Yale University Press, 1996; Alois Nagler. Theatre Festivals of the Medici, 1539-1637. London and New Haven: Da Capo, 1964. 70; Aby Warburg. “I costumi teatrali per gli Intermezzi del 1589: I disegni di Bernardo Buontalenti e il ‚Libro di conti‘ di Emilio de’ Cavalieri.” Gesammelte Schriften. Leipzig: Teubner, 1932. 259 and 394. Strictly speaking, this already applies to the intermezzi produced in 1586. However the 1589 stagings are far better documented. They are also distinguished by greater efficiency. Cf. Cesare Molinari. Le nozze degli dèi. Un saggio sul grande spettacolo italiano nel seicento. Rome: Bulzoni, 1968. 15f.
Telescope, Theater, and the Instrumental Revelation of New Worlds
73
loosely connected the intermezzi was the all-embracing effect of music on men and gods. As the humanists learned from Pythagoras, the universe was numerically and proportionally organized. In music man was able to perceive this divine order. The skies opened for the portrayal of the harmony of the spheres, “and the splendor of what was shown here was so grand…that it really could seem as if paradise had opened and the whole apparatus had become a paradise.”37 Of course, it was the monarchs who guaranteed this harmony. The theatrical apotheosis was meant for them. The complex stage technique itself, however, was more significant than this allegorical reference. In it the prospering of science and the arts under the Medici reign took on a tangible form. The complex scene changes took place with the curtain up. Clouds and planets moved through the room. Mountains appeared and disappeared. Pierids, who had lost in competition against the Muses, were transformed into magpies. A galley with forty men maneuvered on stage, and Arion, who had been thrown overboard, was rescued by a dolphin. All of this required perfectly running stage machinery, impeccable timing, and an army of excellently coordinated stagehands. The actors really only played a very minor role compared to this stage machinery and were always secondary to its workings.38 A description of a similar performance in 1607 suggests that the view behind the scenes alone was fascinating. It was just as delightful to see the winches and elaborate artifacts above the machines; the enormously thick cables, the ropes and cords with which the machines were moved and operated; and the great number of stagehands that were needed to operate them. Every man was in his place and at a given sign the machinery was pulled up, pulled down, moved, or held in a certain position. More than three hundred workers were hired and had to be instructed, a task that required experience, practice, and skill as well as imaginativeness and discernment. It is important to watch out for unexpected incidents as a single spark can ruin everything. It was truly amazing that no accident happened. Just in case, there were special guards who stood by with containers, pots, and buckets filled with water.39 37
38 39
“E tale fu lo splendore, che vi si vide per entró . . . che potette ben parere ad ognuno, che ’l Paradiso s’aprisse, e che Paradiso fosse divenuto tutto l’Apparto (sic!) e la Prospettiva.” Quoted in Mario Fabri and Elvira Zorzi, eds. Il luogo teatrale a Firenze. Milan: Electa ed., 1975. 113. Molinari describes this effect applied to the 1586 Intermezzi. Molinari. Le nozze degli dèi. 15ff. “Gustai non meno vedere sopra le machine gli artifici grandi e gli argani, le gomene grossissime, e le funi, et le corde con che mouano, e maneggiano quelle machine, e ’l numero grandissimò di huomini à maneggiarle, e ciascùno al luogo suo, e ad un cenno calare, alzare e mouere star fremò; pià di trecento huomini à maneg-
74
Florian Nelle
As this astonished description shows, the stage machinery with its technical perfection becomes the allegory of the ideal society that the monarchs had been trying to create. However, it is more than just a mere allegory. It is certainly not by chance that, in 1589, the city’s fortification engineer was given the task of supervising the technical execution of the spectacle. He took careful notes on how Buontalenti organized the stagehands who had to move the clouds, lower the sirens, open the skies, and take care of the oil lamps – just to name a few of their tasks.40 The Medicis’ strict, bureaucratically organized administration worked just as smoothly as the hundred or so stagehands who handled the heavy winches backstage, who made mountains appear and disappear, and who changed the scene completely in a matter of seconds. This medium was the real reflection of the allegories of the Golden Age. It resembled a microcosm of the state, whose brilliant scenery was operated by innumerable, efficiently organized workers and craftsmen.41 It was the most privileged place, where the imaginary was transformed into reality and where fictional worlds from literature, legends, and poetry were rendered into the microcosm of state machinery. After all, its principle was one of constant change. It showed that reality could not only be arbitrarily changed with the aid of ingenious machines. Reality was under the perfect control of the stage engineer and the particular monarch who was to be glorified. This theatrical machine was in perfect control and just like the phantasmagorical intermezzi it was not subject to the course of time.42 This picture was also an example of the Renaissance court itself. In 1651, Thomas Hobbes reduced the theatrum mundi to the machine. He described the state as a perfectly working mechanism, as a machine which consists of countless individuals who play their roles.43 He thus draws
40 41 42
43
giare; si che vi vuole esperieza, eßercito, è prattica, e non men destrezza a ,che ingegno, e giudicio, di auertita auertenza à disordini improuisi, & accidenti in riparare, e prouedere, ché vna fauilla di fuoco può rouinàre ogni cosa. Fù maràuiglia certo, che non vi accadesse disastro alcuno, che vi sono le guardei particolari per questo, con gran vassi, e catini di acque, caldare, e paroli adogni bisogno preparati; Il tutto in somma passò benissimo, sì il tutto fu bene ordinato & essequito; e durò quèsta festa ancora sino alle cinquè hore di notte.” Federico Zuccari. Il passagio per l’Italia con la dimora di Parma. Bologna, 1608. 27. Cf. Warburg. Gesammelte Schriften. 431; Nagler. Theatre Festivals. 78. Cf. Saslow. Medici Wedding. 15. The Renaissance court, however, tended to become a perfectly controllable machine itself. Cf. Adami “L’ingegnere-scenografo.” 163; Carlo Ossola. Dal “Cortegiano” all’ “Uomo di mondo.” Storia di un libro e di un modello sociale. Torino, 1987. 134. Thomas Hobbes. Leviathan. London: Dent, 1934. 1 and 83f.
Telescope, Theater, and the Instrumental Revelation of New Worlds
75
the logial conclusion from the alliance of courtly representation and the theater of machines as it had developed in sixteenth century Italy. Louis XIV perfected this model with his celebrations and divertissements in Versailles. Here the monarch was completely transformed into the magician. In the universe of castle and gardens he had command over space and time, as well as the audience’s astonishment. As André Félibien describes: Once a celebration is properly set up, “there is no one who does not believe that everything happens because of a miracle, that is how amazed one is. In a matter of moments, and without one noticing it, theaters are built, bushes are decorated, and fountains, statues, and banquets are set up, even though this seems possible only by devoting a great deal of time and the efforts of countless workers to it.”44 Here, the surprise effect of changing the scenes moves from the theater of machines into the gardens. It makes the monarch’s apparently unlimited power visible. The instrumental revelation of new worlds does not only happen in science. New Nature, spectacularly staged, is paralleled by a new image of the state in the theater of machines. This image finds its exemplary embodiment in the theatrical machine itself. Despite all the obvious differences between instruments of science and politics, one point can be made: In the seventeenth century, the instrumentalization of science develops in parallel with that of politics. An essential part of this process is the transformation of the instrument into an object of fascination. Both scientific instruments and the theater of machines undoubtedly have a utopian dimension. In the case of theater, however, it is not really utopian. Rather, it is a utopia that has been turned into an exemplary reality. Unlike optical instruments and the new media, the theater of machines is the first in a history of spectacular power institutions, which attempt to establish literary, aesthetic, and social visions in a material reality. Translation: Laura Bohn, Martin Wittenberg
44
“Il n’y a personne qui ne croie que tout s’y fait par miracle tant on est surpris de voir en un moment, et sans qu’on s’en aperçoive, des théâtres élévés, des bocages ornés et enrichis de fontaines et de figures, des collations dressées et mille autres choses qui semblent ne pouvoir se faire qu’avec un long temps et dans l’embarras d’un nombre infini d’ouvriers.” André Félibien. Rélation de la fête de Versailles du 1668 – Les Divertissements de Versailles. Paris: Dédale, Maisonneuve et Larose, 1994. 109.
76
Florian Nelle
WORKS CITED Adami, Giuseppe. “L’ingegnere-scenografo e l’ingegnere-venturiero.” Barocke Inszenierung. Ed. Rudolf Preimesberger and Joseph Imorde. Emsdetten and Zurich: Edition Imorde, 1999. 159-85. Anonymous. “The Ballad of Gresham College.” Ed. Dorothy Stimson. Isis 18 (1932): 103-17. Aubignac, François Hédelin Abbé d’. La pratique du théâtre. Munich: Fink, 1971. Bacon, Francis. “New Atlantis.” The Major Works. Ed. Brian Vickers. Oxford: Oxford University Press, 1996. Blumenberg, Hans. Der Prozeß der theoretischen Neugierde. Frankfurt a.M.: Suhrkamp, 1973. Böhme, Hartmut. “The Metaphysics of Phenomena. Telescope and Microscope in the Works of Goethe, Leeuwenhoek, and Hooke.” Collection, Laboratory, Theater. Scenes of Knowledge in 17th Century. Ed. Helmar Schramm, Ludger Schwarte, and Jan Lazardzig. Berlin: de Gruyter, 2005. 355-93. Boyle, Robert. The Works. Ed. Thomas Birch. Hildesheim: Olms, 1966 [facsimile of the edition London, 1772]. Buck, August. “Die Kunst der Verstellung im Zeitalter des Barocks.” Festschrift der Wissenschaftlichen Gesellschaft an der Johann Wolfgang Goethe-Universität Frankfurt a.M. Wiesbaden: Steiner, 1981. 85-103. Corneille, Pierre. Andromède. Tragédie. Ed. Cristian Delmas. Paris: Didier, 1974. Corneille, Thomas. Circè, tragédie, ornée de machines, de changemens de théâtre, & de musique. Paris, 1690. Descartes, René. Œuvres de Descartes. 13 vols. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1969 Descartes, René. “Discourse on the Method.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdach. Cambridge: Cambridge University Press, 1984. Vol. 1, 111-175. Descartes, René. “Meditations on First Philosophy.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdach. Cambridge: Cambridge University Press, 1984. Vol. 2, 1-62. Descartes, René. The World and Other Writings. Ed. Stephen Gaukroger. Cambridge: Cambridge University Press, 1998. Dryden, John. The Works. 20 vols. Ed. Edward Niles Hooker. Berkeley, Los Angeles, and London: University of California Press, 1971. Evelyn, John. The Diary. 6 vols. Oxford: Clarendon Press, 1955. Fabri, Mario and Elvira Zorzi, eds. Il luogo teatrale a Firenze. Milano: Electa ed., 1975. Félibien, André. Rélation de la fête de Versailles du 1668 – Les Divertissements de Versailles. Paris: Dédale, Maisonneuve et Larose, 1994. Fontenelle, Bernard Le Bovier de. Conversations on the Plurality of Worlds. Trans. H.A. Hargreaves. Berkeley and Los Angeles: University of California Press, 1990. Galilei, Galileo. Dialogue Concerning the Two Chief World Systems. Trans. Stillman Drake. Berkeley: University of California Press, 1953. Gilpin, William. Observations on the Highlands of Scotland. Richmond: Richmond Publishing Co., 1973 [facsimile of the edition London, 1789]. Guarino, Raimondo. La Tragedia e le macchine. Andromède di Corneille e Torelli. Rome: Bulzoni, 1982. Hobbes, Thomas. Leviathan. London: Dent, 1934.
Telescope, Theater, and the Instrumental Revelation of New Worlds
77
Holsboer, S. Wilma. L’histoire de la mise en scène dans le théâtre français de 1600 à 1657. Paris: Droz, 1933. Hooke, Robert. Micrographia or some Physiological Descriptions of Minute Bodies made by Magnifying Glasses. New York: Dover, 1961 [facsimile of the edition London, 1665]. Hooke, Robert. The Posthumous Works. New York and London: Johnson, 1969 [facsimile of the edition London, 1705]. Imorde, Joseph. Präsenz und Repräsentanz oder: die Kunst, den Leib Christi auszustellen. Das vierzigstündige Gebet von den Anfängen bis in das Pontifikat Innocenz X. Emsdetten and Berlin: Edition Imorde, 1997. Konersmann, Ralf. Der Schleier des Timanthes. Perspektiven der historischen Semantik. Frankfurt a.M.: Fischer, 1994. La Bruyère, Jean de. Les caractères. Paris, 1976. Molinari, Cesare. Le nozze degli dèi. Un saggio sul grande spettacolo italiano nel seicento. Rome: Bulzoni, 1968. Nagler, Alois. Theatre Festivals of the Medici, 1539-1637. London and New Haven: Da Capo, 1964. Nelle, Florian. “Descartes und der Regenbogen im Wasserglas – Von der beobachteten zur inszenierten Natur.” Theatralität und die Krisen der Repräsentation. Ed. Erika Fischer-Lichte. Stuttgart and Weimar: Metzler, 2001. 392-410. Nelle, Florian. “Im Rausch der Dinge. Poetik des Experiments im 17. Jahrhundert.” Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst. Ed. Helmar Schramm et al. Berlin: Dahlem University Press, 2003. 140-68. Ossola, Carlo. Dal “Cortegiano” all’ “Uomo di mondo.” Storia di un libro e di un modello sociale. Torino: Einaudi, 1987. Saslow, James M. The Medici Wedding of 1589. Florentine Festival as Theatrum Mundi. New Haven and London: Yale University Press, 1996. Schramm, Helmar. Karneval des Denkens. Theatralität im Spiegel philosophischer Texte des 16. und 17. Jahrhunderts. Berlin: Akademie Verlag, 1996. Shergold, N.D. A History of the Spanish Stage from Medieval Times until the End of the Seventeenth Century. Oxford: Clarendon Press, 1967. Strong, Roy. The Renaissance Garden in England. London: Thames and Hudson, 1998. Vitruvius. The Ten Books on Architecture. Trans. Ingrid D. Rowland. Cambridge: Cambridge University Press, 1999. Warburg, Aby. “I costumi teatrali per gli Intermezzi del 1589. I disegni di Bernardo Buontalenti e il ‘Libro di conti’ di Emilio de’ Cavalieri.” Gesammelte Schriften. Leipzig: Teubner, 1932. 259-300 and 394-441. Zuccari, Federico. “Idea de’ pittori, scultori et architetti.” Scritti d’Arte del Cinquecento. Ed. Paola Barocchi. Milan: Ricciardi, 1973. Vol. 2, 2062-118. Zuccari, Federico. Il passaggio per l’Italia con la dimora di Parma. Bologna, 1608.
FRANK FEHRENBACH
The Pathos of Function: Leonardo’s Technical Drawings
I. A late pen and ink drawing by Leonardo da Vinci, now in Windsor Castle (W 12698 recto, ca. 1513), reverses ideas common in the Renaissance about the sometimes antagonistic relationship between nature and culture (fig. 1). In contrast to his almost contemporary “Deluge Series,” the drawing does not show the worst case scenario of meteorology, destroying the creations of man, but a flood of tools and instruments that fall on the surface of the earth. We identify, among other objects, pincers, hammers, nails, angles, glasses, rulers, rakes, bottles, bagpipes, ladders, clocks, plates, discs, and what is more, among the clouds, a lion. In contrast to a line on the top of the sheet – “Here Adam and here Eve” – which probably belonged to another sketch on the same sheet, later cut off,1 the legend of the actual drawing leaves no doubt about its pessimistic meaning: “O miseria umana, di quante cose per danari ti fai servo. – O human misery, to how many things are you enslaved just for the money.” Related to the lion in the clouds, the sketch becomes a prodigy (cf. Peter Apian, Practica for 1532, Landshut 1531; fig. 2), perhaps alluding to the proverbial technophilia of the Florentines – city of the Marzocco – or a crypto-signature of Leone-Leonardo himself.2 But it is more reasonable, I believe, to interpret the lion as a reference to Pope Leo X, elected in 1513, a person who raised many hopes in Leonardo, which turned, however, into disappointment and complaints. As is well known, Leonardo was denounced before the Pope because of his ana1 2
Cf. Kenneth Clark. The Drawings of Leonardo da Vinci in the Collection of H. M. The Queen at Windsor Castle. 3 vols. London: Phaidon, 1968. Vol. 1, sub num. Cf. Pierce Dominic Britton. “Lionizing Leonardo. A Physiognomic Conceit in Vasari’s ‘Vite.’” Source 22.4 (2003): 10-15; Cecilio Paniagua. “Notes on a Drawing by Leonardo da Vinci.” International Review of Psychoanalysis 13 (1986): 445-52.
The Pathos of Function
79
Fig. 1: Leonardo da Vinci. Deluge of tools. Pen and black chalk, ca. 1510-15. Windsor, Royal Library, No. 12698 recto.
tomical research.3 The drawing could reasonably be interpreted as an allegory of the utilitarian limitations of technology “under the sign of Leo.” However, Leonardo’s bitter lemma emphasizes that greed is the true reason for the deluge of tools that cover the whole surface of the earth. Greed is the driving force behind the emotional involvement in discovery and measurement, ironically documented by countless drawings and manuscript pages of our master technologist himself.4 3
4
On Leonardo’s complaint, cf. Codex Atlanticus, fol. 500. Cf. Domenico Laurenza. “Leonardo nella Roma di Leone X (c. 1513-16).” XLIII Lettura Vinciana. Florence: Giunti, 2004. Cf. also Leonardo’s cryptic contemporary notice: “li medjci mi crearono e desstrussono” (Codex Atlanticus, fol. 429r; on this passage, cf. Carlo Pedretti. The Literary Works of Leonardo da Vinci. Commentary. 2 vols. London: Phaidon, 1977. Vol. 2, 313f., with doubts about a reference to the Florentine family). On the identification of lion and Leo X and on the relevant Medici iconology, cf. Suzanne B. Butters. The Triumph of Vulcan. Sculptor’s Tools, Porphyry, and the Prince in Ducal Florence. 2 vols. Florence: Olschki, 1996. Vol. 1, 58f. On Leonardo’s lion automat for the Florentines in Lyon, on the occasion of the entrance of François I (12 July 1515), cf. Pedretti. Literary Works. 303. On a similar representation by Maarten van Heemskerck, showing the globe covered by instruments and tools (1572), cf. Horst Bredekamp. “Der Mensch als ‘zweiter Gott.’ Motive der Wiederkehr eines kunsttheoretischen Topos im Zeitalter der
Frank Fehrenbach
80
Fig. 2: Peter Apian. Practica for 1532 (Landshut, 1531).
Evidently, the tools that measure, order, straighten, regulate, and adjust are themselves in need of some kind of order. The history of images of technical instruments shows that these trophies of science are marked by a secret unruliness and abundance. Under the command of the new paradigm – mechanics, allegorically boasting as a lansquenet in the period of the Thirty Years’ War5 – this should come to an end. The wellequipped laboratory is not only the place where the world is logically
5
Bildsimulation.” Interface 1. Elektronische Medien und künstlerische Kreativität. Ed. Klaus Peter Dencker. Hamburg: Verlag Hans-Bredow-Institut für Rundfunk und Fernsehen, 1992. 134-47. On the engraving, cf. Ilja M. Veldman and Ger Luijten. Maarten van Heemskerck. (= The New Hollstein. Dutch & Flemish Etchings, Engravings, and Woodcuts 1450-1700). 2 vols. Amsterdam: Sound & Vision Interactive, 1994. Vol. 2, No. 501/1. Cf. the engraving ‘Mechanica and her daughters and sons’ in Joseph Furttenbach’s Mechanische Reißladen . . . Augsburg, 1644. Cf. Jutta Bacher. “Das Theatrum machinarum. Eine Schaubühne zwischen Nutzen und Vergnügen.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 255-97, and her “‘Ingenium vires superat.’ Die Emanzipation der Mechanik und ihr Verhältnis zu Ars, Scientia und Philosophia.” Ibid. 519-55.
The Pathos of Function
81
and numerically regulated; the laboratory itself must tame the instruments, the objects of intervention. The classification of the world goes hand in hand with the spatial disposition of classifying instruments. When dominion over the tools weakens, an ancient chaos reemerges – the epoch post technologiam. Leonardo’s drawing, a writing on the wall about the “second nature” realized by human technology,6 contains, perhaps, the palette of a painter but not – that would have been hard to represent on the tiny sketch – one of the smallest but nevertheless most sophisticated tools of human dominion over nature – the draftsman’s instrument. It is due to the pen that Leonardo’s drawing exists, a rather significant difference to the abundant variety of the not at all heavenly instruments in the representation. It is exactly this peculiarity that identifies the pen with the origin of all tools. The scientific success story of early modern times would have been impossible without the minute but decisive transfers between mental and material concepts realized by the drawing pen. To be sure, this was already evident to the pioneers of this tool. Leonardo is very explicit about this.7 Drawing is nothing but moving a point; not a craft but the gesture of a “light hand,” hardly leaving the trace of bodily activity behind – la pittura è mentale.8 Nevertheless, human culture depends on an almost non-dimensional feather point. “The beginning of painting is the point, followed by the line, the third is the surface.”9 Precisely at this point, the editors’ proposal to insert Leonardo into a book on the culture of instruments in the seventeenth century makes sense. For Leonardo, without pittura all the artes would be impossible, since their codifications require visual signs. Painting – and that means, in this place, drawing, since color is not mentioned at all – not only provides letters for language, numbers for arithmetic, and figures for geometry; it also teaches perspectivists, astronomers, machinists, and engineers (questa insegna alli prospettivi et astrolaghi et alli machinatori et ingegneri).10 Drawing, therefore, is not only a medium conveying knowledge; it also appears as the original invention par excellence, 6
7 8 9 10
“Gravity, force, together with material movement and percussion are the four accidental powers by which the human species in its adorable and various actions appears like a second nature. Because all the visual works of mortals have their being and their death because of these powers.” (Leonardo, Codex Arundel, fol. 151v. ca.1495-97). Cf. Leonardo da Vinci. Manuscript E, fol. 34v. Leonardo da Vinci. Libro di Pittura. Ed. Carlo Pedretti. Florence: Giunti, 1995, § 31c. Ibid. § 3. Ibid. § 23, cf. ibid. § 31b. – A good collection of drawing tools is provided by Maja Hambly. Drawing Instruments. 1580-1980. London: Wilson, 1988.
82
Frank Fehrenbach
giving birth to all other human inventions. Consequently, it would be obsolete to insert painting into the system of the artes liberales. Pittura claims a special rank, comparable, in a different context, only to theology. Obviously, the theoretical claim that visual signs not only communicate, but generate knowledge, points to important epistemological decisions, harking back to the late medieval career of the Aristotelian phantasma. Per imaginem ad ideam: This reveals a clear succession, especially if the inner images are related incontrovertibly to their origin, the senses.11 However, the praise of the minute, of speed, of the almost non-dimensional qualities of point and line, pen and drawing is related to older, Platonic paradigms. For instance, in Alberti’s and Leonardo’s apologia for the eye and the point, what is imprinted by the pen is itself a nusquam, leaving a trace when moved – the line.12 The material medium of the drawing evaporates, and becomes a transitional being between mind and matter, analogous to the spiritus. Later on, this idea could be transformed into an argument against the senses and in favor of art as a domain of the mind, as Georges Didi-Huberman has demonstrated. Vasari states: Drawing, . . . proceeding from the intellect [procedendo dall'intelletto], derives from many things a universal judgment [cava di molte cose un giudizio universale]: as it were a form or idea of all the things in nature [simile a una forma overo idea di tutte le cose della natura], which is exceedingly regular in its proportions. Thus it is that drawing, not only in the bodies of humans and animals, but also in plants, buildings, sculptures, and paintings, recognizes the proportion of the whole to its parts and of the parts to one another and to the whole [conosce la proporzione che ha il tutto con le parti, e che hanno le parti infra loro e col tutto insieme].13
This is the foundation of “composition,” deduced from the living body, since the mutual, proportional usefulness of parts and the whole is, according to Aristotle, a distinguishing formal property of the organism.14 11 12
13
14
Cf. Frank Fehrenbach. Licht und Wasser. Zur Dynamik naturphilosophischer Leitbilder im Werk Leonardo da Vincis. Tübingen: Wasmuth, 1997. 17ff. Cf. Leon Battista Alberti. De pictura 1,2: “Puncta quidem si continenter in ordine iungantur lineam extendent. Erit itaque apud nos linea signum cuius longitudo sane in partes dividi possit. Sed erit usque adeo latitudine tenuissima ut nusquam findi queat.” Giorgio Vasari. Le vite de più eccellenti pittori, scultori et architettori . . . 3 vols. Florence: Gunti, 1568. Vol. 1, 168f. Quoted in Georges Didi-Huberman. Confronting Images. Questioning the Ends of a certain History of Art. Pennsylvania: Pennsylvania State University Press, 2005. 78. Cf. Aristotle. De partibus animalium/Parts of animals. Greek/Eng., transl. A.L. Peck. Cambridge: Cambridge University Press, 1983. 645bff. On the complex history of early modern ‘composition,’ cf. Hans Körner. Auf der Suche nach der ‘wah-
The Pathos of Function
83
Some years later, in Federico Zuccari’s writings, the fusion of idea and disegno will be complete, depriving disegno even of the thin line of the fragile pen and of the emanations at its tip. “Almost another divinity, another productive nature, in which live the things produced by art . . . Internal and external light of the intellect . . . Nourishment and life of all science and practice.”15 But the anti-naturalistic temptations lurking behind this statement are bypassed in the specific cultural situation of the later sixteenth century. Art together with visuality becomes, instead, a “meta-techne”16 of a scientific practice that encourages experiments, proceeding primarily along the paradigms of mapping and collecting, emphasizing at the same time the non-numerical, ‘morphological’ areas of research of the sixteenth century. These are, among others, mainly biology, medicine, alchemy, hydrology, and geology, which feature prominently in Bacon’s Atlantis and remain, together with perspective and optics, the preferred disciplines of many non-Cartesian scientists in the seventeenth and eighteenth centuries. Describing this process of take-over by drawing in 1658, Gassendi coins the accurate phrase, Baconian in spirit (Syntagma philosophicum): “We investigate the objects of nature the same way in which we investigate the things we have created ourselves.”17 But the unity of interpretation and intervention, of knowing and producing can be related, as Wolfgang Krohn recently demonstrated, to Alberti’s efforts to establish the scientific status of architecture.18 This argument brings drawing as a medium of experiment into play. But the drawing’s interventionist, constructivist character remains connected to
15
16 17
18
ren Einheit.’ Ganzheitsvorstellungen in der französischen Malerei und Kunstliteratur vom mittleren 17. bis zum mittleren 19. Jahrhundert. Munich: Fink, 1988; Thomas Puttfarken. The Discovery of Pictorial Composition. Theories of Visual Order in Painting 1400-1800. New Haven and London: Yale University Press, 2000; Frank Fehrenbach. “Komposition.” Metzler Lexikon Kunstwissenschaft. Ideen, Methoden, Begriffe. Ed. Ulrich Pfisterer. Stuttgart and Weimar: Metzler, 2003. 178-83. This text is reprinted in Romano Alberti. Origini e progresso dell’Accademia del Disegno de’Pittori, scultori et architetti di Roma. Padua, 1604. Quoted in DidiHuberman. Confronting Images. 84. Cf. Robert Williams. Art, Theory, and Culture in Sixteenth-Century Italy. From Techne to Metatechne. Cambridge: Cambridge University Press, 1997. Pierre Gassendi. Syntagma philosophicum. Lugduni, 1658. Quoted in Hans Holländer. “Spielformen der Mathesis universalis.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 325-345 (328). Wolfgang Krohn. “Technik, Kunst und Wissenschaft. Die Idee einer konstruktiven Naturwissenschaft des Schönen bei Leon Battista Alberti.” Leonardo da Vinci. Natur im Übergang. Beiträge zu Kunst, Wissenschaft und Technik. Ed. Frank Fehrenbach. Munich: Fink, 2002. 37-56.
84
Frank Fehrenbach
visual evidence and thus to the sensus communis of its addressees and, in the case of technical concepts, to their successful realization. Against this background, favorable to images, experiments, and construction, the social revaluation of the draftsman became almost inevitable. The thirteen-year-old Dürer represented himself in the famous silver point drawing in the Albertina (1484) as a second, teaching Christ Child, with the pointing right hand of the draftsman, fixing his concentrated gaze on an unknown object.19 Even less than the colors of the painter, who mocks, with Leonardo, the dirt-covered, sweating sculptor, drawing left no ugly traces of material action.20 With this, an older controversy was repeated within the guild of the painters themselves – the strife between dirty and ‘philosophical’ activity.21 Even color-orientated princes of painting were better off not to look like fools to the party of (Florentine) intellectuals – and ostentatiously presented the drawing pen rather than the brush, like, for instance, Titian in a lost self-portrait, documented in a woodcut by Giovanni Britto (1550).22 Prior to this, Baldassare Castiglione, in his authoritative concept of the courtly uomo universale, praised drawing as an indispensable activity of nobility.23 The references by Pliny the Elder and Alberti to painting knights and even emperors of antiquity had their effect.24 This development is 19
20 21
22
23
24
Joseph Leo Koerner mentions the parallels with the pointing Jesus of the ‘Holy Family’ in Berlin, ca. 1492-93 (Kupferstichkabinett, SMPK). Cf. Joseph Leo Koerner. The Moment of Self-Portraiture in German Renaissance Art. Chicago: University of Chicago Press, 1993. 14 and 42ff. Leonardo. Libri di Pittura. § 36. On the tradition of the criterion, cf. Claire J. Farago. Leonardo da Vinci’s ‘Paragone.’ A Critical Interpretation with a New Edition of the Text of the ‘Codex Urbinas.’ Leiden: Brill, 1992. 139ff.; also George Ovitt. The Restoration of Perfection. Labor and Technology in Medieval Culture. New Brunswick and London: Rutgers University Press, 1987. Particularly instructive is David Summers. The Judgment of Sense. Renaissance Naturalism and the Rise of Aesthetics. 2nd ed. Cambridge: Cambridge University Press, 1990. 259ff. Joanna Woods-Marsden. Renaissance Self-Portraiture. The Visual Construction of Identity and the Social Status of the Artist. New Haven and London: Yale University Press, 1998. 228f. “Un’altra cosa, la quale io . . . penso che dal nostro cortegiano per alcun modo non debba esser lasciata addietro: e questo è il saper disegnare ed aver cognizion dell’arte propria del dipingere.” Baldassare Castiglione. Il libro del cortegiano. Ed. Walter Barberis. Turin: Einaudi, 1998. XLIX. Gaius Pliny the Elder. Natural History. Trans. H. Rackham. Cambridge: Harvard University Press, 1979; idem. Historia naturalis/Naturkunde. Latin/German. Ed. and trans. Roderich König. Munich and Darmstadt: Artemis & Winkler, 1973ff. Vol. XXXV. 20-22; Leon Battista Alberti. On Painting. New York: Penguin, 1991. 65-66.
The Pathos of Function
85
Fig. 3: Jacopo Pontormo. Alessandro de‘ Medici. Oil on wood panel, 1534. Philadelphia Museum of Art, Johnson Collection.
convincingly mirrored in Jacopo Pontormo’s portrait of the first Florentine Medici Duke, Alessandro, who was assassinated shortly after the completion of the painting (fig. 3). Politically a figure who did not avoid dirty labor, and far from fastidious in matters of love, he appears, on the panel, as a drawing aesthete. On the other side of the picture surface, the painter Pontormo, however, is virtually transformed into Alessandro’s lover, Taddea Malaspina, drawn in this everlasting moment, with a subtle metal point, by the princely bastard – no eternal adoration, but an eternal imitation in the service of the eternal preservation of beautiful bodily features.25 The chiasmus of the gaze transforms the lady who was intended to receive the painting into the portrait of a portrait. Evidently, the mourning of the galantuomo is not only caused by the death of an important protector and distant relative, Pope Clement VII de’ Medici (1534), but also by the implicit absence of the beloved. The im25
Cf. Philippe Costamagna. Pontormo. Milan: Electa, 1994. Cat. no. 72.
86
Frank Fehrenbach
age redeems the prince despite the distance from his lover, like the daughter of a potter at the origins of drawing and/or sculpture,26 in grey contour lines on paper, but colorful in his heart – just as he himself appears, absent, but in colors, before his lover. Even before meeting the sovereign as wood-turner, we encounter the princely draftsman, who fixes his world on paper and who is trained in the tool of access to and subjection of the world through the geometrical magic of perspective.27 Complaints from the subordinates were not missing, since instrumental techniques are not only an analogy of princely dominion, they also set free the delightful passion associated with dilettantism. In situations in which the existence of the community is at risk, the prince should renounce “riding his horse on the turf, carving, turning, painting, or practising alchemy and other things entirely inappropriate to his rank,” as the physician and ducal counselor Caspar Dornau admonished at the beginning of the Thirty Years’ War.28 The sequence of his list interprets painting as a close relative to a speculative and magical art. We should not underestimate the pathos of this imitative recreation of the world through painting/drawing. In Enea Vico’s representation of Baccio Bandinelli’s Accademia (ca. 1550), the more or less drowsy students are busy, at night and by the light of fire and candles, drawing fragments of antique sculpture as well as parts of skeletons – a situation reminiscent of an al26
27
28
Cf. Pliny the Elder. Natural History. Vol. XXXV, 15 and 151. – Significantly, Leonardo conceives the first painting to have been a result of the shadow cast by the sun. Cf. Leonardo da Vinci. Manuscript A, fol. 97v. Cf. for instance the letter of Jacopo de’ Barbari to Frederick the Wise (1500/01): “And without knowledge of the seven Liberal Arts, no convincing painting can be produced by the painters; they have to master these Arts, first geometry, then arithmetic, both indispensable conditions for the measurements of proportion . . . However, to represent these sciences in a painting, philosophy is again required – namely the remarks of Aristotle on the soul, where he discusses how images (species) reach the eye and how the objects are to be arranged on the empty surface of the painting, according to the knowledge of light rays . . . There is no lack of talents, but of leisure and nobility! Because in the age of Alexander the Great, only noble and wealthy men were allowed to practice the art of painting.” Cf. Ulrich Pfisterer, ed. Die Kunstliteratur der italienischen Renaissance. Eine Geschichte in Quellen. Stuttgart: Reclam, 2002. 268f. On the transformations of perspective in the sixteenth century, cf. the survey of Martin Kemp. The Science of Art. Optical Themes in Western Art from Brunelleschi to Seurat. New Haven and London: Yale University Press, 1990. 92. On the history of dilettantism, Wolfgang Kemp. ‘. . . einen wahrhaft bildenden Zeichenunterricht überall einzuführen.’ Zeichnen und Zeichenunterricht der Laien 1500-1870. Ein Handbuch. Frankfurt a.M.: Syndikat, 1979. Cf. also Klaus Maurice. Der drechselnde Souverän. Materialien zu einer fürstlichen Maschinenkunst. Zurich: Ineichen, 1985. Cf. Maurice. Der drechselnde Souverän. 23.
The Pathos of Function
87
chemical laboratory in the service of the revivification of fragmented, fleshless bodies.29 II. The sheer quantity of Leonardo’s surviving technical drawings is terrifying – about six thousand single sheets and manuscript pages, the largest legacy in the history of Renaissance technology. Evidently, many sketches were preserved because they benefited from the aura of the artist’s ingenium – significantly, Vasari compares them to relics.30 What would have happened to Leonardo’s technical drawings if his audience admired him not as the painter of the Milanese Last Supper and the unfinished Battle of Anghiari, but rather as a mechanical genius? The lion’s share of these drawings consists of four subject matters: hydraulics, military technology, flying machines, and the countless sheets dedicated to fundamental problems of theoretical mechanics. The sheer number of drawings demonstrates that Leonardo was unwilling ever to put the pen down; he must have drawn continuously and – significantly – looked repeatedly at his own, older drawings. The range of graphic documents is enormous and can be inscribed ideally into a triangle, bounded by record, phantasmagoric sketch, and diagram. The drawings are in the service of clarification and therefore belong principally to a rhetorical ambience that harnesses the enargeia of the visual. In Leonardo’s work, drawing for the first time claims priority in the history of technology and of the genre of illustrated treatises.31 In Leonardo’s career, the technological ingenium is also set free in social terms. His position is not strictly related to specific commissions and assignments. As a mobile counselor and inventor he moves with astonishing freedom between the different territories – Milan, Florence, Rome, at the end in Amboise and, almost, on the Bosphorus. His salary 29
30 31
On the engraving and on Bandinelli’s academy, cf. Leonard Barkan. Unearthing the Past. Archeology and Aesthetics in the Making of Renaissance Culture. New Haven and London: Yale University Press, 1999. 289ff. Barkan does not notice the representation’s dialectics of fragment and integrity, dead and alive, dark and bright. Cf. Giorgio Vasari. Le vite . . . 6 vols. Ed. Rosanna Bettarini and Paola Barocchi. Florence: Sansoni, 1966-1987. Vol. 4, 28. For a contemporary parallel, Francesco di Giorgio Martini, cf. Paolo Galluzzi, ed. Prima di Leonardo. Cultura delle macchine a Siena nel Rinascimento. Milan: Electa, 1991 (with various examples of technical illustrations without accompanying text, for instance in the circle of Mariano di Iacopo, called Il Taccola). Cf. also Herbert Maschat. Leonardo da Vinci und die Technik der Renaissance. Munich: Profil, 1989; Domenico Laurenza. Le macchine di Leonardo. Florence: Giunti, 2005.
88
Frank Fehrenbach
as a mechanical genius is often independent of specific commissions.32 The unique freedom of Leonardo, however, required the support of elevated rhetoric. Different sources describe Leonardo’s extraordinary eloquence, probably comparable to his graphic skills.33 The famous letter of application (ca. 1480)34 to the then most powerful Duke in Italy, Lodovico Sforza, documented by a transcript in Leonardo’s manuscripts, is a testament to the licentia of the thirty-year-old inventor. What he generously promises closely follows, both in content and in style, a medieval prophet of technological utopia, the Franciscan Roger Bacon, whom Leonardo will repeatedly refer to later on. Among the secreti miei, dangled as bait by Leonardo, pontoon bridges appear beside war ships, armored cars, mine technology, and diving devices – exactly those technologies of movement on and under earth and water imagined by Roger Bacon two hundred years before, which were intended to renew, as Horst Bredekamp has shown, the technological miracles of antiquity.35 Projects like the sickle car belong to this classicizing context (fig. 4). In this case however, Leonardo warns that this is a terribly powerful device, capable of bringing death and mutilation indifferently to friend and foe. Leonardo’s reputation as a technician, the willingness even of skeptical governments like the Venetian or the Florentine Republics to commission enormous (ultimately failed) public works, Leonardo’s skill at turning pragmatic situations into poetic contexts, must be linked to his unique skills as a draftsman. I would like to highlight the characteristics of his drawings with the following keywords – perspective, geometry, and especially, movement. A drawing in Windsor of ca. 1503/04 (fig. 5), produced in the context of Cesare Borgia’s threats to attack the Florentine Republic, reflects these three criteria in nuce. Leonardo depicts a sophisticated defensive device that can attack an adversary who has already surmounted the outer ring of a fortress from behind. Leonardo demonstrates the bombardment by hidden underground ordnances as an actual event, drawing the semicircular trajectories of the projectiles in perfect beauty, almost like the abstract play of fountains. The graphic display of the foreshortened, elliptical 32 33 34 35
Cf. Paolo Galluzzi. Gli ingegneri del Rinascimento da Brunelleschi a Leonardo da Vinci. Florence: Giunti, 2001. 47ff. “E con le parole volgeva al sì e al no ogni indurata intenzione.” Vasari. Le vite. Ed. Bettarini and Barocchi. Vol. 4, 37. Leonardo da Vinci. Codex Atlanticus, fol. 1082. Cf. Domenico Laurenza. “Leonardo. Le macchine volanti.” Le macchine del Rinascimento. Ed. Giovanni Morello. Rome: Retablo, 2000. 145-87; Horst Bredekamp. Antikensehnsucht und Maschinenglauben. Die Geschichte der Kunstkammer und die Zukunft der Kunstgeschichte. Berlin: Wagenbach, 1993.
The Pathos of Function
89
Fig. 4: Leonardo da Vinci. Sickle Car. Pen and ink, ca. 1490. Turin, Biblioteca Reale, No. 15583.
Fig. 5: Leonardo da Vinci. Military Project. Pen and ink, ca. 1503/04. Windsor, Royal Library, No. 12275.
90
Frank Fehrenbach
curves is extraordinary. Leonardo attached great importance to the fact that the distance between the projectiles on the ground was exactly identical. This signals that the defense technology participates in the geometrical perfection of the perspectival grid. The event takes place without the anecdotal addition of attacking or dead soldiers, on a perfectly plain, clean, and evenly lit terrain. The viewer observes an ideal process, a programmed sequence. In order to achieve greater visibility, the fortification is sliced by a sharp cut. Furthermore, Leonardo depicted the perspective of neither the invaders nor the defenders, but rather presented an ideal oblique view, floating over the walls and out of reach of the projectiles. Later called the “military perspective,” this view implies a general standing upon an imaginary command post on a hill. Leonardo’s drawing, which anticipates the esthetics of the video war, seems to utter the words: “You, who look at me, you can only be a strategist; you have the perspectiva.” The military rhetoric of perspective – the analogy of ‘looking through’ and shooting through – can be witnessed in a later illustration in Agostino Ramelli’s Diverse et artificiose machine (1588) (fig. 6). Only a draftsman like Leonardo, who masters perspective and at the same time the anatomy of the moving body, is able to maximize the privilege of the viewer. Leonardo’s image of a cannon foundry was probably produced in the 1490s (fig. 7), in Milan, at the time the most important center of the military industry in Europe. The enormous complexity of the drawing should not divert us from the fact that some of the activities remain rather unclear, especially the extremely strenuous operations of the workers on the right. The main subject is the assembly of a newly cast giant cannon on a cart. Many naked workers are busy with different, but synchronized actions that require the utmost physical effort. The drawing is especially persuasive in two respects. First, it contrasts the single teams, unable to overlook the whole scene, with the viewer who does; second, it presents the workers as functioning without a coordinating foreman, almost like an automated operational process. The viewer is implicitly transformed into the overseer or the programmer of the actions. More than that, he is the only one able to survey the entire working process, from the casting in the foreground with the movement of the cannon on pulleys through the application of hoists, to the later, implicit loading of the ordnance with the balls piled in the background. The viewer also supervises the well-equipped arsenal of cannon barrels (protected by a roof) and other materials distributed over the area. While the hardware – weapons, buildings, tools – is clearly outlined and well defined by perspective, the naked workers appear to be ephemeral collectives, an
The Pathos of Function
91
Fig. 6: Agostino Ramelli. Diverse et artificiose machine (Paris, 1588, fig. CXLVI).
organic apparatus, or rather transient incarnations of the forces of traction and thrust, with quite permeable contours. The combination of perspective order and sketchiness, dynamic and organic forms, brings us to another point – the calculated non-finito of technical drawings, the rhetoric of the draft. Here too, Leonardo was a pioneer. Although there must have been many drawings like the representation of a big dredger digging a river-bed (fig. 8) – illustrations which by their perfection make the personal style of the draftsman almost disappear – the majority of sketches which contributed to the fame of Leonardo certainly displayed the evidence of their production. This can be deduced not only from the surviving technical drawings. It is evident because Leonardo’s artistic production is also characterized by an aversion to completion and by a continuous preference for exploratory, open
92
Frank Fehrenbach
Fig. 7: Leonardo da Vinci. Cannon Foundry. Pen and ink, ca. 1495. Windsor, Royal Library, No. 12647.
techniques, with well-known disastrous results.36 It must have been an integral part of the technological rhetoric of Leonardo to reveal his process of sketching and drawing. The secretary of the Cardinal of Aragon, who visited Leonardo in 1517 in Clos Lucé, reported that Leonardo proudly presented the long series of his manuscripts – and we know how ‘chaotic’ most of these manuscripts appear.37 36
37
Cf. (for Leonardo’s early work) Michael Wiemers. Bildform und Werkgenese. Studien zur zeichnerischen Bildvorbereitung in der italienischen Malerei zwischen 1450 und 1490. Munich and Berlin: Deutscher Kunstverlag, 1996. 265ff. After his visit to Leonardo in Cloux (10 October 1517), Antonio de’ Beatis notes: “Ha anche composto de la natura de l’acqua, de diverse machine et altre cose [note the sequence!], secondo ha riferito lui, infinità di volumi et tutti in lingua volgare.” Antonio de Beatis. Die Reise des Kardinals Luigi d’Aragona durch Deutschland, die Niederlande, Frankreich und Oberitalien 1517 bis 1518. Ed. Ludwig von Pastor. Freiburg: Herdersche Verlagshandlung, 1905. 143; cf. Carlo Pedretti. Leonardo da Vinci on Painting. A Lost Book (Libro A). Berkeley and Los Angeles: University of California Press, 1964. 109, and Pedretti. Literary Works. Vol. 2, 140.
The Pathos of Function
93
Fig. 8: Leonardo da Vinci. Project of a Dredger. Pen and ink, ca. 1500. Codex Atlanticus, fol. 4 recto. Milan, Biblioteca Ambrosiana.
I suspect that the ‘ingenious’ state of Leonardo’s technical projects contributed enormously to his persuasive eros. In 1504, Leonardo supported the plan of the Florentine Republic – a favorite project of his friend, Niccolò Machiavelli – to divert the course of the Arno before it reached Pisa, in order to cut off running water and the sea from the arch-enemy.38 Soon after work began, the operations at the construction site, where over two thousand workers were employed at some time, had to be abandoned in face of the insurmountable difficulties of this giant enterprise. At the same time, Leonardo developed a project to render the Arno navigable from Florence to the sea. The entire Val di Chiana, to the south of Arezzo, was to be flooded, serving as a reservoir for the river, and the rapids near Empoli were to be bypassed by a broad diversion of the Arno to the north. Part of the project was a canalised tunnel at Serravalle (near Lucca), after which the river would return to its original bed.39 Leonardo’s faith in the production of almost unlim38
39
On this project, cf. Nicolai Rubinstein. “Machiavelli and the Decoration of the Hall of the Great Council in the Palazzo Vecchio.” Musagetes. Festschrift für Wolfram Prinz. Ed. Ronald G. Kecks. Berlin: Mann, 1991. 275-85. Ladislao Reti. “Leonardo the Technologist. The Problem of the Prime Mover.” Idem and Bern Dibner. Leonardo da Vinci, Technologist. Norwalk: Burndy Library, 1969. 90.
94
Frank Fehrenbach
Fig. 9: Leonardo da Vinci. Map of the Arno. Various media, 1503/04. Windsor, Royal Library, No. 12279.
ited forces was, at that time, still intact, although the development of the flying machine sounded a note of caution.40 In 1495, Leonardo contemplated a cannon that would somehow shoot the earth off its central position in the universe – a technological phantasmagoria, significantly connecting Archimedes, gunpowder, and Nicolaus Copernicus.41 A large map of 1503/4 (fig. 9), also in Windsor Castle, is a very elaborate product in different media (pen with brown ink, black chalk, bistre), showing the new course of the Arno in a bold, sweeping line. In its combination of geographical exactitude (note the tributaries of the Arno!) and “impure,” alternative channel courses, drawn nevertheless with great decisiveness, the drawing is a telling example of the ideal 40
41
Cf. Frank Fehrenbach. “Hier stehe ich, aber mein Auge durcheilt die Räume. Die Vogelstudien und Flugversuche von Leonardo da Vinci.” Frankfurter Allgemeine Zeitung (April 27, 2002): 47. Leonardo da Vinci. Codex Madrid I, first flyleaf recto: “Se possibile fussi fare una bonbarda, che ’l mondo fussi sua ballotta, e che sicome una bonbarda gitta una balotta d’un braccio 3 miglia, che si pò misurare il tal corso 9000 braccia, cioè 9 mila ballotte. Noi possiamo adunque dire, che tal bonbarda gitterebbe il nostro mondo novemila volte la grandeza del diamitro d’esso mondo distante da ssè. [Sare]bono a settemila miglia per mondo, sarebono 63 migliara di miglia.”
The Pathos of Function
95
mixture (and its ethical connotations) of know-how, accuracy, and courage – scientia, diligentia/cura, and potestas audendi – which must have hypnotized administrators. The combined trust in God and in the good outcome of high tech projects seem to mesmerize not only modern governments; as is well known, the plans for the new Florentine Cathedral, for instance, were revised in the second half of the 14th century without clear ideas about how to close the giant space of the cupola.42 Leonardo’s ingenious drawing of the Arno sheds some light on the significant moments of these planning procedures. We can imagine Leonardo drawing the new course of the river on the large sheet of paper, over the landscape, in a sweeping left-hand gesture, mimetically following the course of the water from right to left, with a wet brush, and perhaps right before the eyes of an admiring audience. III. Before 1500, Leonardo as draftsman used mainly metal point (a medium not allowing corrections) and, to an even larger extent, pens of different size. As a common medium for sketching and writing, the pen underlines the priority of ‘painting’ over writing; even letters are, in this perspective, nothing but another form of drawing.43 After 1500, Leonardo employs a wider range of graphic materials – red and black chalks, pastels, charcoals, and watercolors.44 These diverse media create different visual effects, and these effects are, in turn, related to the particular physical qualities of the objects. The early drawings of Leonardo already document a very accurate differentiation of media. Quite often the draftsman works with the dry stilus, creating colorless, ‘ghostlike’ lines imprinted in the paper. These lines are visible only with close attention and oblique illumination; they cannot be reproduced by any means. The reproductions we hold in our hands, in more and more weighty publications, show just the ‘surface’ of an œuvre that borders on the invisible. After drawing these ‘ghost lines,’ the subtle figurations of the silver point adumbrate the represented object. The later application of pen and ink defines contours and shadows – or lead to new deformations by means of dynamic pentimenti. 42 43 44
Cf. Margaret Haines. “Brunelleschi and Bureaucracy: The Tradition of Public Patronage at the Florentine Cathedral.” I Tatti Studies 3 (1989): 89-125. Cf. Leonardo. Libro di pittura. § 23. Cf. Francis Ames-Lewis. La matita nera nella pratica di disegno di Leonardo da Vinci (= Lettura Vinciana, vol. 41). Florence: Giunti, 2002.
96
Frank Fehrenbach
After 1500, Leonardo’s graphic media reflect the qualities of the objects represented. This is particularly evident, for example, in the large studies of facial expressions for the Battle of Anghiari, intended to decorate the republican sanctuary of Leonardo’s city, the Sala del Maggior Consiglio in the Palazzo Vecchio. Two drawings in Budapest depict older, experienced warriors and a youthful soldier (fig. 10 & 11). While the wrinkled soldiers, with their wrathful grimaces, were executed in the darker and more brittle black chalk, for the shouting young man Leonardo chose the harder, sharper and at the same time ‘blood-colored’ red chalk (Italian: sanguigna).45 A few years later, Leonardo worked on a series of topographical representations of the Alps (W 12414), employing red chalk again, this time on red paper, with some white highlights, in order to visualize the firm consistency of the mountains. The more or less contemporary series of fantastical mountain landscapes, showing wavering, overhanging, and exploding rocks, was executed, however, in soft charcoal. Adhering only loosely to the paper ground, the medium complements the fragility and transient nature of the subject. Material disconnection instead of firm joints features prominently in the so-called ‘Deluge Series’ (W 12376-12386, ca. 1515), for which Leonardo initially, experimented with pen and ink. When he realized the inadequacy of the medium he completed the other drawings in black chalk. The mimesis, or – more adequately – metonymy of the media present the material qualities of the objects of representation. A similar metonymy can be observed in the development of Leonardo’s technique of hatching. As a strongly conventionalized and regionally astonishingly constant workshop technique, hatching has never been a subject of systematic representational analysis.46 It was not a topic of literary debate; therefore historical hermeneutics is not provided with a terminological framework. This lack of interest is astonishing, since the graphic representation of three-dimensionality proves the inadequacy of the com45
46
Domenico Laurenza. “Corpus mobile. Ansätze einer Pathognomik bei Leonardo.” Leonardo da Vinci. Natur im Übergang. Beiträge zu Kunst, Wissenschaft und Technik. Ed. Frank Fehrenbach. Munich: Fink, 2002. 257-301. Cf. Bernhard Degenhart. “Zur Graphologie der Handzeichnung. Die Strichbildung als stetige Erscheinung innerhalb der italienischen Kunstkreise.” Kunstgeschichtliches Jahrbuch der Bibliotheca Hertziana 1 (1937): 223-343. (I would like to thank Heiko Damm for bringing this article to my attention.) – In the recent, historically very broad and theoretically ambitious study of drawings by David Rosand, hatching is not discussed systematically (Drawing Acts. Studies in Graphic Expression and Representation. Cambridge: Cambridge University Press, 2002). Convincing remarks on the tactility of hatching types, ibid. 107 and 110f. (Leonardo); on temporality: 111 (Leonardo) and 206f. (Michelangelo).
The Pathos of Function
97
Fig. 10: Leonardo da Vinci. Study for the “Battle of Anghiari.” Black chalk, ca. 1504. Budapest, Szépmüvészeti Múzeum, No. 1774.
Fig. 11: Leonardo da Vinci. Study for the “Battle of Anghiari.” Red chalk, ca. 1504. Budapest, Szépmüvészeti Múzeum, No. 1775.
98
Frank Fehrenbach
Fig. 12: Leonardo da Vinci. Study of a Head. Silver point, ca. 1485. Turin, Biblioteca Reale, No. 15572.
mon distinctions between model and mimesis. Graphic translations of objects are imitations and constructions at the same time.47 The system of lines and hatches that make objects visible do not at all disappear behind the illusion of the object; they remain present and emphasize the artificiality of the representation. As a young artist, Leonardo experimented with different hatching techniques, as can be seen, for instance, in the famous landscape drawing of 1473.48 In Milan, Leonardo elaborated a different style, dominated by extremely accurate, tightly parallel diagonal hatching (fig. 12), giving the impression of a “rain that densely patters, an evenly oblique rain shower”, as Anny Popp observed in 192849 – a rain that at the same time visualizes a ‘ground’ (of space, of the paper) from which the rep47
48 49
On the non-mimetic and non-conventional character of graphic elements, cf. Meyer Schapiro. “On Some Problems in the Semiotics of Visual Art. Field and Vehicle in Image-Signs.” Theory and Philosophy of Art. Style, Artist, and Society. New York: Braziller, 1994. 1-32 (esp. 27f.). Florence, Uffizi, GDS inv. no. 8P. Anny E. Popp. Leonardo da Vinci. Zeichnungen. Munich: Piper, 1928. 25.
The Pathos of Function
99
resented bodies emanate in subtle gradation. Applied by a virtuoso like Leonardo, the volume of objects seems to be defined as a relief, in very fine layers, in order to suggest their plasticity. The technique is, to be sure, closely related to contemporary engravings.50 However, this graphic practice tends to ‘freeze’ depictions of action and movement. The most prominent stylistic modification in Leonardo’s entire œuvre concerns the elaboration of hatching with curved lines, complementing and substituting the straight, parallel diagonal lines of the earlier drawings. This style can be observed, for the first time, in the earlier of the two Madrid Codices, rediscovered only in 1966. Completed around 1495, this is the most elaborate of all of Leonardo’s manuscripts, dedicated entirely to the axioms of mechanics. Leonardo describes the theoretical foundations of mechanical forces and presents basic mechanical elements (lever, spring, spindle, etc.), an alphabet of mechanics that precedes every application in concrete machines. With this, Leonardo moves, historically for the first time, from engineering to systematic research in mechanics, or from technique to technology.51 With significant delay, books on machines from the end of the sixteenth century onwards have followed Leonardo’s model (e.g. Salomon de Caus). In Codex Madrid I, Leonardo frequently hatches various elements of the machines with curved lines.52 On fol. 45r (fig. 13), for instance, Leonardo draws a spring intended for a clockwork. The body of the spring, defined by curved lines, appears as a monumental form, full of inner plastic energy. Its function – the slow, circular release of stored forces – is expressed in an almost physiognomic manner by the method of hatching. It seems as if Leonardo discovered the adequate graphic means for Vitruvius’s famous defintion of the machina: “Machina est continens e materia coniunctio, maximas ad onerum motus habens virtutes. Ea movetur ex arte circulorum rotundationibus quam Græci țȣțȜȚțȘȞ țȚȞȘıȚȞ appellant.”53 In Leonardo’s mechanical alphabet the aforementioned persuasive esthetical intentions reappear: perspective, geometry, 50
51 52
53
Cf. Konrad Oberhuber, Jay A. Levenson, and Jacquelyn L. Sheehan. Early Italian Engravings from the National Gallery of Art. Washington: National Gallery of Art, 1973. XVff. Galluzzi. Prima di Leonardo. 47ff. Cf. Pietro C. Marani. “Leonardo dalla scienza all’arte; un cambiamento di stile, gli antefatti, una cronologia.” Fra Rinascimento, Manierismo e Realtà. Scritti di storia dell’arte in memoria di Anna Maria Brizio. 2nd ed. Ed. idem. Florence: Barbèra, 1984. 41-52. Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard University Press, 1962. Vol. 2, 274 [my emphasis].
100
Frank Fehrenbach
Fig. 13: Leonardo da Vinci. Spring. Pen and ink, ca. 1492. Codex Madrid I, fol. 45r. Madrid, Biblioteca Nacional.
and movement. Even the individual elements of the mechanical organism seem ‘rhetorical.’ Consequently, the perceiving eye of the viewer merges with the modeling hand of the artist-technician. In his early Manuscript A (ca. 1492), Leonardo observed that the eyes, while looking at moving water, cannot remain immobile.54 In his late Manuscript E (1513/14), he adds that seeing is normally in motion, thus drawing a line through the visual field – to extrapolate, like a draftsman who moves a point in creating a line.55 This theoretical background is made manifest by means of curved 54 55
“Se tu riguardi il movimento dell’acqua, l’occhio tuo non si può fermare ma fa a similitudine delle cose vedute.” Manuscript A, fol. 58v. Cf. Manuscript E, fol. 80v; also 34v, 35r (in the context of transformation geometry; cf. Matilde Macagno. Geometry in Motion in the Manuscripts of Leonardo da Vinci. Iowa City: University of Iowa Press, 1987). On the relationship of curved hatching and the mobile eye, cf. also Donald S. Strong. Leonardo on the Eye. An English Translation and Critical Commentary of MS. D in the Bibliothèque Natio-
The Pathos of Function
101
hatching. The eye itself weaves the objects of images, and its movement projects a modeling veil on moving things, merging with their graphic creation and expansive dynamics.56 Again, Leonardo’s graphic strategy connects the privileged viewer with the production of images. His technical drawings, too, connect the producing instrument – the pen of the draftsman – to the eye of the viewer. To put it somewhat pointedly: The graphic suggestion works, because it is the addressee himself who produces it. Every spectator becomes a ‘drawing sovereign.’ IV. The rhetoric of drawing at the origins of instruments and technical practices – the later history of this constellation in the sixteenth and seventeenth centuries would be a separate and urgent research project, necessarily including not only the printed, but also the manuscript sources of the history of technology in a comprehensive way.57 At this point, only a few concluding observations could be made. The printed engineering treatises of the sixteenth and seventeenth centuries address mainly the socially high-ranking dilettante and belong therefore to the field of lusus. The rare and highly appreciated dedicated machine book manuscripts of the sixteenth century present their (partly non-functional) mechanical devices in perfect integrity, frequently relying in their perspective constructions on Francesco di Giorgio. In 1671, in a letter to the Duke of Hanover, the later King George I of England, Gottfried Wilhelm Leibniz rejects these ‘dead’ – immobile – machines, created for a distanced, to extrapolate: esthetic contemplation.58 At this time, Leonardo’s singular ‘physiognomic’ skills in presenting machines and
56
57 58
nale, Paris, with Studies on Leonardo’s Methodology and Theories of Optics. New York: Garland, 1979. 408f. But compare the too narrow view, following Gotthold Ephraim Lessing, of Daniela Lamberini: “It is impossible to represent movement on a two-dimensional surface (and dynamic motion is the intrinsic characteristic of the machina, as opposed to the fabrica, or building, the static machine par excellence).” Daniela Lamberini. “Machines in Perspective. Technical Drawings in Unpublished Treatises and Notebooks of the Italian Renaissance.” The Treatise on Perspective. Published and Unpublished (= Studies in the History of Art 59 – Symposium Papers XXXVI). Ed. Lyle Massey. New Haven and London: Yale University Press, 2003. 213-33 [214]. Cf. the survey by Hélène Vérin. La gloire des ingénieurs. L’intelligence technique du XVIe au XVIIIe siècle. Paris: Albin Michel, 1993. Cf. Lamberini. “Machines in Perspective.” 213 (following Manlio Brusatin. Storia delle linee. Turin: Einaudi, 1993. 52).
102
Frank Fehrenbach
their elements in a dynamic, rhetorical manner were already substituted by graphic conventions and by the increasingly important three-dimensional model making.59 Historically, the hand-‘writing’ of the drawing technician survived only in the stenographic milieu of the workshop manuscripts and in architectural drawing.60 At the end of the eighteenth century, the military engineer Gaspard Monge and his collaborator Jean Nicolas Pierre Hachette ‘purged’ the genre of technical drawings of its last rhetorical elements (among them the perspectival relationship between viewer and functional object). What remains are the elements of machines, ordered in table form, and visualized as section and diagram.61
WORKS CITED Alberti, Leon Battista. On Painting. New York: Penguin, 1991. Ames-Lewis, Francis. La matita nera nella pratica di disegno di Leonardo da Vinci (= Lettura Vinciana, vol. 41). Florence: Giunti, 2002. Aristotle. De partibus animalium/Parts of animals. Trans. A.L. Peck. Cambridge: Cambridge University Press, 1983. Bacher, Jutta. “Das Theatrum machinarum. Eine Schaubühne zwischen Nutzen und Vergnügen.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 509-18. Bacher, Jutta. “‘Ingenium vires superat.’ Die Emanzipation der Mechanik und ihr Verhältnis zu Ars, Scientia und Philosophia.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 519-55. Barkan, Leonard. Unearthing the Past. Archeology and Aesthetics in the Making of Renaissance Culture. New Haven and London: Yale University Press, 1999. Bredekamp, Horst. “Der Mensch als ‘zweiter Gott.’ Motive der Wiederkehr eines kunsttheoretischen Topos im Zeitalter der Bildsimulation.” Interface 1. Elektronische Medien und künstlerische Kreativität. Ed. Klaus Peter Dencker. Hamburg: Verlag Hans-Bredow-Institut für Rundfunk und Fernsehen, 1992. 134-47. Bredekamp, Horst. Antikensehnsucht und Maschinenglauben. Die Geschichte der Kunstkammer und die Zukunft der Kunstgeschichte. Berlin: Wagenbach, 1993. Britton, Pierce Dominic. “Lionizing Leonardo. A Physiognomic Conceit in Vasari’s ‘Vite.’” Source 22.4 (2003): 10-15. Brusatin, Manlio. Storia delle linee. Turin: Einaudi, 1993. 59 60 61
On the relationship of drawing and model, cf. also Massimo Scolari. “Elementi per una storia dell’assonometria.” Casabella 550 (March 1984): 42-49. On architectural drawings, cf. Werner Oechslin. “Von Piranesi zu Libeskind. Erklären mit Zeichnung.” Daidalos 1 (1981): 15-19. Cf. Sandrina Khaled. “Pikturale Graphismen der Technik, 1569-1870.” Bilder in Prozessen (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1.1). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2003. 64-78.
The Pathos of Function
103
Butters, Suzanne B. The Triumph of Vulcan. Sculptor’s Tools, Porphyry, and the Prince in Ducal Florence. 2 vols. Florence: Olschki, 1996. Castiglione, Baldassare. Il libro del cortegiano. Ed. Walter Barberis. Turin: Einaudi, 1998. Clark, Kenneth. The Drawings of Leonardo da Vinci in the Collection of H. M. The Queen at Windsor Castle. 3 vols. London: Phaidon, 1968. Costamagna, Philippe. Pontormo. Milan: Electa, 1994. Degenhart, Bernhard. “Zur Graphologie der Handzeichnung. Die Strichbildung als stetige Erscheinung innerhalb der italienischen Kunstkreise.” Kunstgeschichtliches Jahrbuch der Bibliotheca Hertziana 1 (1937): 223-343. Didi-Huberman, Georges. Confronting Images. Questioning the Ends of a Certain History of Art. Pennsylvania: Pennsylvania State University Press, 2005. Farago, Claire J. Leonardo da Vinci’s ‘Paragone.’ A Critical Interpretation with a New Edition of the Text of the ‘Codex Urbinas.’ Leiden: Brill, 1992. Fehrenbach, Frank. Licht und Wasser. Zur Dynamik naturphilosophischer Leitbilder im Werk Leonardo da Vincis. Tübingen: Wasmuth, 1997. Fehrenbach, Frank. “Hier stehe ich, aber mein Auge durcheilt die Räume. Die Vogelstudien und Flugversuche von Leonardo da Vinci.” Frankfurter Allgemeine Zeitung (April 27, 2002): 47. Fehrenbach, Frank. “Komposition.” Metzler Lexikon Kunstwissenschaft. Ideen, Methoden, Begriffe. Ed. Ulrich Pfisterer. Stuttgart and Weimar: Metzler, 2003. 178-83. Furttenbach, Joseph. Mechanische Reißladen . . . Augsburg, 1644. Galluzzi, Paolo, ed. Prima di Leonardo. Cultura delle macchine a Siena nel Rinascimento. Milan: Electa, 1991. Galluzzi, Paolo. Gli ingegneri del Rinascimento da Brunelleschi a Leonardo da Vinci. Florence: Giunti, 2001. Haines, Margaret. “Brunelleschi and Bureaucracy: The Tradition of Public Patronage at the Florentine Cathedral.” I Tatti Studies 3 (1989): 89-125. Hambly, Maja. Drawing Instruments. 1580-1980. London: Wilson, 1988. Holländer, Hans. “Spielformen der Mathesis universalis.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 325-345. Kemp, Martin. The Science of Art. Optical Themes in Western Art from Brunelleschi to Seurat. New Haven and London: Yale University Press, 1990. Kemp, Wolfgang. ‘. . . einen wahrhaft bildenden Zeichenunterricht überall einzuführen.’ Zeichnen und Zeichenunterricht der Laien 1500-1870. Ein Handbuch. Frankfurt a.M.: Syndikat, 1979. Khaled, Sandrina. “Pikturale Graphismen der Technik, 1569-1870.” Bilder in Prozessen (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1.1). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2003. 64-78. Körner, Hans. Auf der Suche nach der ‘wahren Einheit.’ Ganzheitsvorstellungen in der französischen Malerei und Kunstliteratur vom mittleren 17. bis zum mittleren 19. Jahrhundert. Munich: Fink, 1988. Koerner, Joseph Leo. The Moment of Self-Portraiture in German Renaissance Art. Chicago: University of Chicago Press, 1993. Krohn, Wolfgang. “Technik, Kunst und Wissenschaft. Die Idee einer konstruktiven Naturwissenschaft des Schönen bei Leon Battista Alberti.” Leonardo da Vinci. Natur im Übergang. Beiträge zu Kunst, Wissenschaft und Technik. Ed. Frank Fehrenbach. Munich: Fink, 2002. 37-56. Lamberini, Daniela. “Machines in Perspective. Technical Drawings in Unpublished
104
Frank Fehrenbach
Treatises and Notebooks of the Italian Renaissance.” The Treatise on Perspective. Published and Unpublished (= Studies in the History of Art 59 – Symposium Papers XXXVI). Ed. Lyle Massey. New Haven and London: Yale University Press, 2003. 213-33. Laurenza, Domenico. “Leonardo. Le macchine volanti.” Le macchine del Rinascimento. Ed. Giovanni Morello. Rome: Retablo, 2000. 145-87. Laurenza, Domenico. “Corpus mobile. Ansätze einer Pathognomik bei Leonardo.” Leonardo da Vinci. Natur im Übergang. Beiträge zu Kunst, Wissenschaft und Technik. Ed. Frank Fehrenbach. Munich: Fink, 2002. 257-301. Laurenza, Domenico. “Leonardo nella Roma di Leone X (c. 1513-16).” XLIII Lettura Vinciana. Florence: Giunti, 2004. Laurenza, Domenico. Le macchine di Leonardo. Florence: Giunti, 2005. Leonardo da Vinci. Il Codice Atlantico di Leonardo da Vinci nella Biblioteca Ambrosiana di Milano. Ed. Accademia dei Lincei. Transcribed by Augusto Marinoni. 24 vols. Florence: Barbèra, 1973-1980. Leonardo da Vinci. Codices Madrid. Ed. Ladislao Reti and Augusto Marinoni. 5 vols. Frankfurt a.M.: Fischer, 1974. Leonardo da Vinci. Libro di Pittura. Ed. Carlo Pedretti. Florence: Giunti, 1995. Leonardo da Vinci. Il codice Arundel 263 nella British Library. Ed. Carlo Pedretti. Transcribed with commentary by Carlo Vecce. Florence: Giunti, 1998. Macagno, Matilde. Geometry in Motion in the Manuscripts of Leonardo da Vinci. Iowa City: University of Iowa Press, 1987. Marani, Pietro C. “Leonardo dalla scienza all’arte; un cambiamento di stile, gli antefatti, una cronologia.” Fra Rinascimento, Manierismo e Realtà. Scritti di storia dell’arte in memoria di Anna Maria Brizio. 2nd ed. Ed. idem. Florence: Barbèra, 1984. 41-52. Maschat, Herbert. Leonardo da Vinci und die Technik der Renaissance. Munich: Profil, 1989. Maurice, Klaus. Der drechselnde Souverän. Materialien zu einer fürstlichen Maschinenkunst. Zurich: Ineichen, 1985. Oberhuber, Konrad, Jay A. Levenson, and Jacquelyn L. Sheehan. Early Italian Engravings from the National Gallery of Art. Washington: National Gallery of Art, 1973. Oechslin, Werner. “Von Piranesi zu Libeskind. Erklären mit Zeichnung.” Daidalos 1 (1981): 15-19. Ovitt, George. The Restoration of Perfection. Labor and Technology in Medieval Culture. New Brunswick and London: Rutgers University Press, 1987. Paniagua, Cecilio. “Notes on a Drawing by Leonardo da Vinci.” International Review of Psychoanalysis 13 (1986): 445-52. Pastor, Ludwig von, ed. Antonio de Beatis. Die Reise des Kardinals Luigi d’Aragona durch Deutschland, die Niederlande, Frankreich und Oberitalien 1517 bis 1518. Freiburg: Herdersche Verlagshandlung, 1905. Pedretti, Carlo. Leonardo da Vinci on Painting. A Lost Book (Libro A). Berkeley and Los Angeles: University of California Press, 1964. Pedretti, Carlo. The Literary Works of Leonardo da Vinci. Commentary. 2 vols. London: Phaidon, 1977. Pfisterer, Ulrich, ed. Die Kunstliteratur der italienischen Renaissance. Eine Geschichte in Quellen. Stuttgart: Reclam, 2002. Pliny the Elder, Gaius. Historia naturalis/Naturkunde. Latin/German. Ed. and trans. Roderich König. Munich and Darmstadt: Artemis & Winkler, 1973ff. Pliny the Elder, Gaius. Natural History. Trans. H. Rackham. Cambridge: Harvard University Press, 1979.
The Pathos of Function
105
Popp, Anny E. Leonardo da Vinci. Zeichnungen. Munich: Piper, 1928. Puttfarken, Thomas. The Discovery of Pictorial Composition. Theories of Visual Order in Painting 1400-1800. New Haven and London: Yale University Press, 2000. Reti, Ladislao. “Leonardo the Technologist. The Problem of the Prime Mover.” Idem and Bern Dibner. Leonardo da Vinci, Technologist. Norwalk: Burndy Library, 1969. Rosand, David. Drawing Acts. Studies in Graphic Expression and Representation. Cambridge: Cambridge University Press, 2002. Rubinstein, Nicolai. “Machiavelli and the Decoration of the Hall of the Great Council in the Palazzo Vecchio.” Musagetes. Festschrift für Wolfram Prinz. Ed. Ronald G. Kecks. Berlin: Mann, 1991. 275-85. Schapiro, Meyer. “On Some Problems in the Semiotics of Visual Art. Field and Vehicle in Image-Signs.” Theory and Philosophy of Art. Style, Artist, and Society. New York: Braziller, 1994. 1-32. Scolari, Massimo. “Elementi per una storia dell’assonometria.” Casabella 550 (March 1984): 42-49. Strong, Donald S. Leonardo on the Eye. An English Translation and Critical Commentary of MS. D in the Bibliothèque Nationale, Paris, with Studies on Leonardo’s Methodology and Theories of Optics. New York: Garland, 1979. Summers, David. The Judgment of Sense. Renaissance Naturalism and the Rise of Aesthetics. 2nd ed. Cambridge: Cambridge University Press, 1990. Vasari, Giorgio. Le vite . . . 6 vols. Ed. Rosanna Bettarini and Paola Barocchi. Florence: Sansoni, 1966-1987. Veldman, Ilja M. and Ger Luijten. Maarten van Heemskerck. (= The New Hollstein. Dutch & Flemish Etchings, Engravings, and Woodcuts 1450-1700). 2 vols. Amsterdam: Sound & Vision Interactive, 1994. Vérin, Hélène. La gloire des ingénieurs. L’intelligence technique du XVIe au XVIIIe siècle. Paris: Albin Michel, 1993. Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard University Press, 1962. Wiemers, Michael. Bildform und Werkgenese. Studien zur zeichnerischen Bildvorbereitung in der italienischen Malerei zwischen 1450 und 1490. Munich and Berlin: Deutscher Kunstverlag, 1996. Williams, Robert. Art, Theory, and Culture in Sixteenth-Century Italy. From Techne to Metatechne. Cambridge: Cambridge University Press, 1997. Woods-Marsden, Joanna. Renaissance Self-Portraiture. The Visual Construction of Identity and the Social Status of the Artist. New Haven and London: Yale University Press, 1998.
NICOLA SUTHOR
“Il pennello artificioso”: On the Intelligence of the Brushstroke Artificem instrumenta, dominum qualis sit domus ostendere. Pomponius Gauricus, 1504
The various ways in which the brush has been used historically, its role in several painting traditions, has by no means received enough attention among the important questions of art history. This omission should not be overlooked any further, because it allows us to avoid developing the concept of art outside the realm of practice, a predominant tenet of pre-modern academic art theory.1 As a mechanical link between hand, paint, and canvas, the brush embodies a technique that its ostensibly ‘organic’ execution conceals – as the age-old saying goes in art: ars est celare artem (it is art to conceal the art). The assimilation of the technique of the brush in the corporality of the hand produces the desired ‘naturalness’ of the painterly mark, which then has nothing ‘artificial’ left in it. The mastery of the hand according to this concept of art is that it demonstrates that it is possible to forget about technique as an unnatural extension of the hand. The disregard of the brush in art history validates this conception of painting. But how we then explain, that in the art theory of the seventeenth century the brush becomes the instrument of art par excellence (and not just technique) and synonymous with the artist himself?2 Instead of simply understanding the brush as an extension of the hand, it is worth investigating more precisely the specialization of the 1
2
In his fundamental study Style in the Art Theory of Early Modern Italy, Philip Sohm states: “Not only is the derivation of ‘maniera’ from ‘mano’ (hand) absent from every definition [of style], but manual aspects pf painting such as technique and brushwork are either suppressed or only discreetly acknowledged. This was not a product of ignorance or oversight so much as a deliberate avoidance of the uncomfortable truth that painting is a physical act.” Philip Sohm. Style in the Art Theory of Early Modern Italy. Cambridge: Cambridge University Press, 2001. 153. This disguising of the resistance of technique while the master is at work is one of the reasons why the materiality of the instrument is not given much attention.
“Il pennello artificioso”
107
Fig. 1: Nicolas Régnier. Self-Portrait with the Portrait of Vincenzo Giustiniani on the Easel. Oil on canvas, 1623-1624.
artistic hand in terms of a distinction between its various possibilities relative to the use of the brush. In the art history of the period, the painting intelligence that arises between hand and instrument, and that fashions its seamless intertwining, finds its manifest expression in the stroke of the brush. In the first section of this investigation, we shall refer to an exemplary self-portrait from the seventeenth century showing the artist at work because it ostensibly represents the technique of painting. Because it shows the handling of the brush in multiple ways, this self-presentation serves as the background to the second part, which is dedicated to the theorectical examination of the way in which the brush becomes pars pro toto of art in the seventeenth century. The Artist Painting Nicolas Régnier’s Self-Portrait with the Portrait of Vicenzo Giustiniani on the Easel (1623-1624) can illustrate for us a use of the painter’s instrument that is exemplary for its time (fig. 1). The right hand, its sig-
108
Nicola Suthor
nificance underscored by the handling of the light and its intensity, employs the brush with a certain ease. The little finger is raised elegantly. The action of the right hand, which is occupied with the mixing of the colors, is assisted by the work of the left. Hidden under the palette, supporting it, the left hand holds all the implements ready for the right. The cultivated air of the painter, evidenced not least by the elegance of his dress, also emanates from his clearly exemplary effortless handling of his instrument. Next to the palette, which weighs down on the palm of the hand, the left hand grasps the bundle of brushes that are squeezed between the forefinger and thumb. The middle finger pinches the painting cloth against the inside of the hand, while the little finger clasps the maulstick. These physical efforts of the left (through the inverted mirroring of the right!) are hidden in Régnier’s self-portrait beneath the highlighted right hand. The elegance of holding the brush precisely at the moment of a pause in the act of painting, in order to glance at the observer, underscores the gentility of the painter who is dressed in a fine black robe. The brush is held in the middle in order to facilitate a certain ease of execution. Régnier portrays himself in the act of executing the portrait. The facial features represented in the portrait resting on the slightly canted easel are realistically painted and appear to emerge from the tondo. The consequent shortening of the portrait is practically nullified by the turn of the head. Both the painter and the subject look at us with interest. The black garment allows the painter to recede into the background, his corporality dissolves into the darkenss of the surrounding area, a single fleck of light emphasizes the contours of his shoulder. His subject, the Marquese Vincenzo Giustiniani, on the other hand, is situated much more smoothly in his surroundings. His deep red velvet cloak and the beige fur collar underscore his physical presence. In contrast to the painter, whose facial features and hand are highlighted by a cool light, the likeness of the Marquese is built up from below – from the roughly sketched cloak to the finely detailed features of the face. The moment of interruption in the process of painting has already been proposed by Annick Lemoine as the defining characteristic of the concetto of the painting.3 In fact, at first glance the viewer of the 3
“Régnier stellt sich in dem Moment dar, als sei er eben bei der Arbeit unterbrochen worden, um auf den Betrachter zu reagieren.” Annick Lemoine, “Nicolas Régnier, zugeschrieben, Porträt des Marquese Vincenzo Giustiniani.” Caravaggio. Die Sammlung Giustiniani und die Berliner Gemäldegalerie in Preussen [Exhibit. cat.]. Ed. Silvia Danesi Squarzina. Milan: Electa, 2001. 200.
“Il pennello artificioso”
109
scene has the impression of having interrupted a meeting. Yet the concentrated gaze of the painter, which is directed toward us, like the action of his right hand – mixing with a thickly bristled brush the paints on the palette – present for us a complex situation, in which the viewer who is addressed can only be the commisioner of the painting, which is Giustiniani himself. The person portrayed in the portrait is the model and at the same time the sovereign observer of the scene. Two gazes, that of the artist, who studies his subject in order to find the correct tone, and that of the sitter, who checks the likeness of the portrait and observes with interest the act of painting, meet in the painting. The selfportrait, which returns the observant gaze of the Marquese, functions in this sense like a mirror. The gaze of the connoisseur, which dominates the scene and is addressed at the level of reception of the painting, is joined with that of the artist, who in the crafting of the painting has already captured his point of view as observer. His painterly skill is focused on the completion of the – nevertheless hidden to us – painting that he has (according to the narrative of the painting: in vivo) in front him. The thin brushes that he holds in his left hand point toward the collar and the as yet incompletely finished red cloak, which has already been layed out in bravura strokes. Shades of light and shadows suggest the material likeness of velvet. The right hand with the genteelly raised little finger conspicously holds the thickest brush in the hand. The paint, which he has mixed under his brush, implies that the painter is ready to put the finishing touches to the fur of the collar. In contrast to the finely modelled face, here the characteristic style of the quick and surely placed brushstroke must not fail mimetically, instead it can even be articulated in its physicality. The Freedom of the Brush The polarization of painting technique between diligenza and prestezza, demonstrated by the two painting styles (in assembling bold and delicate strokes), is a constant of the art theory of the early modern period, thematizing the inspired handling of the brush. The mode of execution of prestezza provides elbowroom for an explicitly unconventional gestural development of the brush as painting instrument. In contrast to diligenza, in which the artist restrains himself and is enslaved by capturing a likeness, prestezza is linked to the notion of artistic freedom. At the end of the eighteenth century, Reverend William Gilpin in his Three Essays: On Picturesque Beauty (1794) succinctly formulates the expression of artis-
110
Nicola Suthor
tic freedom in the brushstroke: “A stroke may be called free when there is no appearance of constraint. It is bold when a part is given for the whole, which it cannot fail to suggest. This is the laconism of genius.”4 In the first systematic reference book of the visual arts, the Dictionnaire by Claude-Henri Watelet and Pierre Charles Lévesque, brush handling is, in a broad range of entries, assigned a central role in the evaluation of ars. According to these entries, the quick and surely placed brushstroke demonstrates “the great freedom of the hand or of the brush” (“la grande liberté de main ou de pinceau”). This “freedom of execution” (“cette liberté d’exécution”) is understood here as an expression of “noble boldness” (“cette noble fierté”) that distinguishes the artist from the commonherd. The refinement of painterly execution is emphasized by virtue of an equation with the aristocracy. Artists can (like the “nobles”) also high-handedly promote the fortune with which they gamble, without being limited by their needs. The mastery stemming from artistic freedom owes itself thus to a penchant for risk. Not diligence measured in working hours, but rather speculation on the prize seems to promise high art. Half the entries for “manœuvre,” a term that refers to the application of paint and the execution of the brush, concern the unconventional application of paint. According to these entries, Rembrandt’s student, Arent de Gelder, applied paint not with the brush but rather with the thumb or the palette knife. In other words, the process of mixing paint was extended from the palette to the canvas itself, and thus a production step was traceable on the actual painting that up to this point had merely been part of a preparatory stage and had never been employed in the execution of the final painting. In the article “Main,” Lévesque writes about the headstrong decision of Cornelis Ketel, who after twenty years of painting like others with the brush, gave it up in order to try painting with his fingers. This “tour de force” was not special enough (assez singulier) for him however: at first he used the left hand instead of the right, then he even attempted to paint with his feet. For the sake of originality, Ketel employed his own physicality in order to circumvent conventional practice and, by means of this virtuoso accomplishment, to exhibit obstinacy. In his Traité de la peinture (1765), Michel-François Dandré Bardon describes Rembrandt’s painting trick that consisted of punching the canvas in order to create a depression in the surface. In this way he achieved a convincing depth in parts of the painting which was not possible by 4
William Gilpin. Three Seáis. On Pitturesque Beauty, etc. (1794). 61f.. Quoted in Edgar Wind. Art and Anarchy. Illinois: Northwestern University Press, 1985. 83.
“Il pennello artificioso”
111
the use of color alone. This example, which Dandré Bardon ultimately dismisses as not worthy of imitation, shows according to him however the effect of the “grands stratagêmes” in creating special color effects.5 Inverting the conventional usage of the painting implement – Gelder employed the brush handle in order to paint lace trimmings – allows the creation of “singular effects” (“des effets singuliers”). The author points out that such an ‘opportune’ extension of conventional painting techniques demonstrates the originality of his paintings, which, for the collector, ultimately comes at a price. In naming the exceptional boldness of the painter and the chance (la fortune) that provokes this “noble fierté” as the very factors that make artistic freedom possible, Watelet cites two criteria that Vasari, a good two hundred years earlier, claimed were an indication of the absence of disegno. Significantly Vasari disqualifies with this fundamental criticism – he regards disegno as the father of all the arts – his Venetian contemporary Tintoretto, whose brushwork is ruled by chance (caso) and boldness (fierezza).6 When Vasari, in his thoroughly critical commentary, adds “almost to demonstrate that this art is a joke (baia),” then he begins to provide the first hints at the identity of the ‘genius artist.’ The description of his painting style as “stravagante, capriccioso, presto e risoluto”7 specify characteristics of a maniera, that do not become evidence of a daring “exécution” until Watelet. Ultimately the art literature of the seventeenth century stylizes Tintoretto as a ‘genius’ trailblazer of prestezza-painting. And just a century after Vasari, Carlo Ridolfi glorifies Tintoretto’s resolute brushwork: “Every stroke of his brush (ogni striscio del suo pennello) was a glorious indication of his immortality (un tratto di gloria per la sua immortalità).”8
5
6
7
8
Francois Dandré Bardon. Traité de Peinture suivi d’un essai sur la sculpture pour servir d’introduction à une histoire universelle, relative à ces beaux-arts. 2 vols. Paris, 1765. Vol. 1, 229. “Ha lavorato a caso e senza disegno, quasi mostrando che quest’arte è una baia.” Giorgio Vasari. Le Vite de’ più eccellenti pittori, scultori e architettori. 6 vols. Ed. Rosanna Bettarini and Paola Barocchi. Florence, 1976. Vol. 5, 468. In his study “Die Kunst des Capriccio,” Roland Kanz corrects the assessment of Vasari’s criticism of Tintoretto, which had formerly been interpreted as disapproval. “Trotz Vasaris mangelnder Bereitschaft, die koloristischen Errungenschaften der venezianischen Malerei zu würdigen, wird Tintoretto hohe Wertschätzung zuteil.” Roland Kanz. Die Kunst des Capriccio. Berlin: Deutscher Kunstverlag, 2002. 218. Carlo Ridolfi. “Vita di Tintoretto.” Le Meraviglie dell’Arte ovvero le vite degli illustri pittori veneti e dello stato. Vol. 2. Padua: Arnaldo Forni Editore, 2002 [1648]. 192.
112
Nicola Suthor
Pars pro toto When Francesco Scannelli in Il Microcosmo della Pittura (1657) states that Correggio keeps his drawings in the tip of his brush,9 he inverts the traditional hierarchy of drawing as preliminary step, to be followed afterwards by painting. Pierre Lebrun, in his Recueil des Essaies des Merveilles de la Peinture (1635) also holds the opinion that the invenzione comes first from the hand and in fact from the ‘maneuver’ of the brush: “Whereas the inspiration for poets comes from their heads, where the poetic nerve lies, the painters have it in their fingertips and the tip of their brushes.”10 According to this, the creative process of painting differs from that of poetry insofar as in the former the spiritual imagination is the heart of invenzione, while the Poesis of painting is intrinsically liberated by the act of painting. Viewing the brush as a tool for the expression of artistic originality leads finally to a merging of maniera and brush. The high esteem of the instrument as pars pro toto is characteristic of the descriptions of art in the seventeenth century. In Bologna, in 1633, an anthology appears that compiles the panegyrics to Guido Reni’s “Rape of Helena,” under the auspicious title Il trionfo del pennello. And in 1666 Luigi Scaramuccia composes his investigation on Le finezze de’pennelli italiani. In this art discourse we find a multitude of similar formulations such as, “i primi esquisiti Penelli, cioè à dire di un Raffaello, di un Parmigianino, Titiano, Giulio Romano.”11 André Félibien takes a critical stance against this merging in his L’Idée du peintre parfait (1707): At times the word brush stands for the source of all the aspects of painting. Hence one says that Raphael’s painting of the Transfiguration is the most beautiful work that his brush created, and sometimes one understands by this the works themselves, and one says for example, that of all the painters of antiquity, the cleverest brush was that of Apelles. But here the word brush only 9
10
11
“Correggio, era solito rispondere, c’haveva i suoi dißegni nella stremità de’Pennelli.” Francesco Scannelli. Il Microcosmo della Pittura. Bologna: Nuova Alfa Editore, 1989 [facsimile of the edition Cesena, 1657]. 359. “Les poètes ont leurs inspirations dans la teste où est la nerve poëtique, et les peintres au fin bout des doigts et à la pointe scarante du pinceau.” Pierre Lebrun. “ Recueil des essaies des merveilles de la peinture de Pierre Lebrun, Peintre (Brussels Manuscript 1635).” Medieval and Renaissance Treatises on the Arts of Painting. Original Texts with English Translations. Ed. Mary P. Merrifield. New York: Dover, 1999. 767. Luigi Scaramuccia. Le finezze de’ pennelli italiani, ammirate, e studiate da Girupeno . . . Pavia, 1674. 177.
“Il pennello artificioso”
113
means the superficial way in which it is handled in order to apply the paint. And when the same colors are not too stirring, too unsettled because of the motion of a heavy hand, but on the contrary are the result of a motion that was free, direct, and easy, then one says that the work comes from a good brush. But the free brush is of little value when it is not driven by the head, and it is only worth familiarizing the one with the other once the painter is in full control of his artistic intelligence.12
The relationship between the hand and the artistic intelligence that drives the action of the painter is complex in that the relationship of dependence can be reversed, so that what arbitrarily comes into being under the hand can in turn provoke the intellect. In his Le finezze de'pennelli, Scaramuccia postulates that the artist – according to the ancient artistic motto: Nulla dies sine linea (never a day without a line) – should train the hand, and with it the intellect (essercitar la mano, e con essa l’intelletto). This alliance of hand-intellect, which is expressed in a sure handling of the brush, can be illustrated by the painting concept of bravura, which took root in the art theory of the seventeenth century. Filippo Baldinucci defines “Bravura” in his Vocabolario toscano dell’arte del disegno (1681) as a “particular boldness, or ferocity of motion, intense in every movement of the figure, in which sometimes a little harshness is noticeable.”13 Corroborating this definition, a fundamental distinction is made in the entry for “De colpi” between the “gran bravura e padronanza del pennello e d’colori” and the paintings that are described as “sfumate, o affaticate” (faded or tired). The signification of this term is derived from the adjective “bravo” that in everyday speech signifies primarily competence and ability – a coinage found in Vasari’s Lives and also present in Pellegrino Antonio Orlandi’s Abecedario pittorico (1753) – can be read as the admiration of a mastery of practice, evidenced by the sure handling of the medium (“bravura e padronanza del pennello e d’colori”). The boldness of the bravura is due to itself to a technical skill that willfully seeks to prove itself – in the sense of acts of bravery – in the impassioned act of painting. Marco Boschini’s La Carta del navegar pittoresco (1664) was groundbreaking for Baldinucci’s coinage. Penned as an expression of praise in Venetian dialect and in verse, this history of Venetian painting glorifies the “great bravura” of the Venetian maniera when alluding to her 12
13
See the entry “Pinceau” in the Dictionnaire des arts de peinture, sculpture et gravure. 5 vols. Ed. Claude-Henri Watelet and Pierre Charles Lévesque. Paris: Prault, 1792. Vol. 5, 61-73. “Brauura f. Vna certa fierezza, o furia di mouimento veemente in ogni operazione della figura, alla quale non disdice alle volte vn poco di durezza.” Filippo Baldinucci. Vocabolario toscano dell’arte del disegno . . . Florence, 1681. 23.
114
Nicola Suthor
“great knowledge, great brushes, and great colors” (gran saver, gran peneli e gran colori!). The arranging of artistic knowledge, implement, and medium in almost one breath ties theory to practice by virtue of the instrument as modus of expression. The fact that artistic knowledge inspires the brushstroke is also echoed by Francesco Algarotti, who in his Saggio sopra la Pittura singles out the “grande bravura del pennello:” “the hand only works when it obeys the intellect.”14 The notion of the influence of the intellect on the artistic hand, as well as the idea of the “spirituality” (Boschini) of the brushstroke, is directed towards the criticism of bravura painting according to which the quickness (prestezza) and spontaneity (prontezza) of prestezza-painting is ‘mind-less,’ since it is executed without thinking and thus appears to be ruled less by the intellect than the daring hand of the artist. The disegno, which precedes the painting and has to be faithfully transformed in it (but which can never correspond to an intellectual “purity” since the reproduction is always deficient) does not guarantee the intellectuality of the painting. It is rather the confident handling of the painting instrument that can set the artistic genius free. Only with practice does the artist attain a savoir-faire15 that distinguishes his brush as unique. When, at the end of the eighteenth century in his treatise Della pittura italiana (1797), Antonio Maria Zanetti commends the painter Jacopo Bassano for having taken his own path in order “to paint with knowledge” (“di dipingere con sapere”) having “put aside that great diligenza and great finitezza to adopt instead bravura;”16 when he calls Tintoretto the “most terrible painter” (“il più terribile pittore”), the “ruler” (“il Padrone, e disponitore”) because he knows how to paint “con fierezza e soma intelligenza;” and when he ultimately singles out “la prontezza, e il sapere” and “la prestezza” of Luca Giordano and notes that “his way of painting with such a free and experienced brush frightens the professori,”17 then it becomes clear that in the course of the eighteenth century 14
15 16
17
“Il primo avvertimento sia questo; che la mano non corra sì che non ubbidisca all’intelletto.” Francesco Algarotti. Saggio sopra la Pittura. Ed. William Spaggiari. Rome: Archivio Guido Izzi, 2000. 57. “Ché colpizar xe l’Arte del saver.” Marco Boschini. La carta del navegar pitoresco. Ed. Anna Pallucchini. Venice: Istituto per la collaborazione culturale, 1966. 68. “Lasciò da parte questi la gran diligenza, e la gran finitezza, e vi pose in suo luogo la bravura.” Antonio Maria Zanetti. Della pittura veneziana. Venice: Albrizzi, 1797. 209. “Tante e sì belle sono le opere, che di questo pittore in Venezia si veggono, che sarebbe Viniziano non non sia, deesi discorrere anche di lui tra i Viniziani. Ebbe egli un particolar suo modo di dipingere con un maneggio del pennello così franco e inteso, che spaventa i professori.” Ibid. 227.
“Il pennello artificioso”
115
Fig. 2: Simon Vouet. Self-Portrait. Oil on canvas, ca. 1620.
the establishment of the term bravura del pennello in art history is connected to a concept of painting that grants forced obstinacy in manual practice a genuinely intelligible value. Simon Vouet: The Artist as a Bravo To illustrate how bravura was thematized in painting before it was coined as a term, we can refer to two exemplary self-portraits by Simon Vouet. Instead of fashioning his art with his instrument (like Régnier), Vouet visualized in both paintings the process of making art in demonstrative brushstrokes. In the earlier self-portrait of 1620 the pronounced turn of the head indicates a spontaneous shift toward the viewer (fig. 2). This pointed moment is rendered in paint by a few ‘prompt’ brush strokes that sketch his dress and which provide as it were the foundation against which the finely modeled facial features stand out. A sophisticated
116
Nicola Suthor
painterly game emerges between illusionary portrayal and simultaneous making the materiality of the medium visible. While Vouet works out the skin of the features here in a highly life-like manner, which is in some places lightly reddened and in others appears to be slightly perspiring (suggested by the points of light on the end of the nose and above the lip), he avoids a realistic depiction of fabric in the clothing. Instead of being draped in appealing softness, the gaze of the spectator is directly hit by the brown primer of the canvas, upon which the garment is crudely sketched with thick brush strokes in a mimetically abbreviated manner. This mode of applying the paint with a very thick brush is rudimentary in that the responsibility of mimetically finishing the painting is almost handed over to the viewer. In the awareness of what is intended representationally he can show himself to be a connoisseur of art and experience the process of painting. This compelling translation into bravura strokes demands the input of the viewer’s power of imagination. The skillfully painted open collar casts a strong shadow on the coat. The prominent padded shoulder profile is given shape solely by means of a broad band of ochre-hued brush strokes. Beneath these, black brush strokes marking the beginning of the sleeves mould a shadowy area underneath the brim. Next to these, the broad black brush strokes of the jacket are sprinkled with white flecks that animate the painting surface. An interesting relationship arises between the fields of light reflection and the tips of the collar. The quickly and directly placed brushstrokes of the sleeve coalesce into a pointed shape, which shows a marked void. The view through the mimetic plane onto the painting ground corresponds to the gaps in the lace. While here the white nevertheless stands for the fabric, the use of black and white on the painting ground indicates bright and dark light and shadows. These un-mimetic, purely color sensations that emerge on the clothing – almost like a gleam – and visually manufacture its plasticity, are physically tangible in the full materiality of the paint. While the sword that underscores his “aristocratic bearings”18 is delineated here somewhat near the edge by a black squiggle and a white highlight, it becomes in a later portrait the central feature of the painting (fig. 3). The elegantly curved ornamental sword consisting of handle, blade, and hilt – painted with precise and carefully placed glimmering highlights –falls out of the picture plane, which is itself painted 18
According to Klingsohr-Leroy the characteristics of the portrait lie in the renouncing of the artist’s status to show instead aristocratic bearing. Cathrin Klingsöhr-Leroy. Das Künstlerbildnis des Grand siècle in Malerei und Graphik. Vom “Noble Peintre” zum “Pictor Doctus.” Munich: Fink, 2002. 34.
“Il pennello artificioso”
117
Fig. 3: Simon Vouet. „Le Spadassin“. Oil on canvas, ca. 1620.
with broad passes of the brush on top of a brown underpainting. While the white lines on the decoratively convoluted handle depict lighting effects, the majority of the coarse brushwork – seen here mostly in the sleeves – does not work mimetically. Rather than focusing on the illusionary power of a single brushstroke, we will examine here the range in styles of brushwork in the foreground that fashion the lighting and paint masses. The variety in the execution of the brush stroke demonstrates the ability of the hand, finely modeled, which is also emphazised in the painting – the handle of the sword almost placed in relationship to it. The correlation between the relaxed pointing (at himself) right hand and the sword highlighted by a few prompt brushmarks can be understood as intentional. The linking of “manus” and “spada” is the seminal idea of bravura painting, demonstrated here avant la lettre. Hence Bo-
118
Nicola Suthor
schini in his La Carta del navegar pitoresco (1660) glorifies Venetian painters as “Bravi” (daring heroes) and “Capitani” who practice “brave acts” (ati sì fieri) in painting.19 He regards Tintoretto as primus inter pares, in “his sword lives life and death.”20 In the Breve Instruzione Boschini extolls his “fulminating” brushes, by whose power he “eclipsed and terrified the greatest masters of art.” Through this bravura he ultimately achieved “the absolute domination of art” (“l’assoluto domino dell’Arte”).21 To speak metaphorically of brushwork as fencing is not arbitrary, instead it has a definite correlation to the life of the artist himself: a switch from sword to brush or vice-versa is frequently mentioned in early modern artists’ biographies. As an example Pietro Maria da Crevalcore was “not any less talented (bravo!) with the brush than he was with a sword in his hand.”22 The development of training the hand by the use of a weapon, as Orlandi reported in his Abecedario pittorico,23 was taken to an extreme in the entry of a particular Spanish painter, Estevan March, who as an example significantly excelled as a painter of battle scenes. Before starting to paint he would lock himself in his room full of weapons, which he would wield with great effort until exhausted, in order then to finally “grab the brush and beautifully give form to the dead, half dead, and wounded.”24
19 20 21
22
23
24
Marco Boschini. La carta. 88. On Tintoretto’s Occupation of Brescia, cf. ibid. 229. “Poiché con il suo fulminante pennello ha colpeggiato così fieramente, che ha fatto arrestare ed atterrire i più generosi Camioni dell’Arte.” Ibid. 730. Concerning this, cf. Massimiliano Rossi. “La peinture guerrière. Artistes et paladins à Venise au XVIIe siècle.” La Jérusalem délivrée du Tasse. Poésie, peinture, musique, ballet. Ed. Giovanni Careri. Paris: Klincksieck, 1999. 67-108. “Pietro Maria da Crevalcore, bravo non meno con pennello, che con la spada alla mano.” Pellegrino Antonio Orlandi. Abecedario pittorico. Venice: Pasquali, 1753. 132, 203, 334. Likewise, in the Vite written by Bernardo di Dominici, the double training as painters as well as fencers is a leitmotiv (e.g. in the vita of Cavaliere Mattia Preti). Bernardo de Dominici. Vite de’ Pittori, Scultori ed Architetti napoletani. Naples, 1742-45. On Lorenzo Gabrieri, called Il Nipote di Carracci: “Addestrata dunque la mano al pennello, alla spada, ed al suono . . . con la spada alla mano bravamente si difese.” Orlandi. Abecedario pittorico. 346. “Si chiudeva nella sua stanza ch’era piena di armi, e quelle con grande fatica a maneggiar si metteva, fino che si stancava, ed allora prendeva i pennelli, ed a maraviglia esprimeva morti, semivivi e feriti.” Ibid. 465. Cf. also a similar entry in Don Antonio Palomino Velasco. Las vidas de los pintores y estatuarios eminentes españoles . . . London, 1724. 75 (entry CIII on Estevan Marc).
“Il pennello artificioso”
119
Ars et métier The mastery of the fierezza (boldness) of the brush stroke develops from the concept of furor pittoresco – an idea that Vasari had already thematized when basing the quickness of painting, which he himself experienced as a painter, on the following argument: Just as the poetry that is brought about by poetic fury is true, good, and better than the poetry that is labored, so is the design work of the artist better when it is made with the strokes of this fury in comparison to when they are belabored with effort and strain.25
Finally, in his essay Dell’Entusiasmo delle belle arti (1769), the Jesuit Saverio Bettinelli explains the “pennello di fuoco” by the elevated state of the passion of the “genius.” Bettinelli sets genius apart from ingenuity (and aligns the genius of warriors with the ingenuity of courtiers)26 when he asserts that a “dominator” does not need to study and is not subject to regulation. The “sovereign geniuses” (“questi Genji sovrani”), whose features are unbridled wildness and excess “flee from laboriousness, and do not tolerate the long run” (“fuggirono la fatica, non suffrirono la lunghezza”). Therefore rapidità (quickness) is considered one of their outstanding characteristics. In his theory of enthusiam, which elaborates the contemporary use of the term, Bettinelli frees it from its literal translation “divine inspiration” in order to look for an affective foundation. As an incarnation of enthusiam, Bettinelli names the warrior, who with a flick of the wrist externalizes an inner aggression.27 This concept of 25
26 27
This notion supports the topos of the tragic painter who could not take his hands from his work. His “troppo diligenza” suffocates the artistic furor from the beginning. “Nelle bozze molte volte, nascendo in un subito dal furore dell’arte, si sprima il suo concetto in pochi colpi, e che per contrario lo stento e la troppo diligenza alcuna fiata toglia la forza et il sapere a coloro che non sanno mai levare le mani dall’opera che fanno. E chi sa che l’arte del disegno, per non dir la pittura solamente, sono alla poersia simili, sa ancora che come le poesie dettate dal furore poetico sono le vere e le buone e megliori che le stentate, così l’opere degli huomini eccellenti nell’arte del disegno sono migliori quando son fatte a un tratto dalla forza di quel furore, che quando si vanno ghiribizzando a poco a poco con istento e con fatica.” Vasari. Le Vite. Vol. 3, 52. Saverio Bettinelli. Dell’Entusiasmo delle belle arti. Milan, 1769. 192. “La parola Entuasiasmo ci vien dal Greco, e significa letteralmente Ispirazione divina, ed in Toscano s’interpreta dall Crusca Sollevamento di mente, Furor divino. Nell’uso poi, e nella commune intelligenza è preso per un agitamento dell’anima, che si palesa ancora nel corpo, e fa l’uomo parlare, ed operare con impegno, con forza, e con valore straordinario in certe occasioni, studj, ed imprese. Nell’abuso di questo termine suol intendersi per un trasporto, una violenza, un’audacia, che tende al male, e all’errore . . . Un uomo vivace, ed un collerico, un tenero cuore, ed ar-
120
Nicola Suthor
‘genius’ – thus developed from life experience – manifests itself in the art literature of the time in a growing awareness of the handling of the media (le faire). Refering to Tintoretto’s painting, the painter Charles-Nicolas Cochin, in his notes on his trip to Italy (1758), underlines “the enthusiam of his genius and the furor of his brushes.”28 When Cochin notes that Tintoretto’s art lacks “fire” in its finishing because his imagination “cools” during the completion,29 then the willful painterly approach is an indication for him of the intuitive skill of the artist that releases itself in the spontaneous act but that cannot be translated or overworked. It is only experienced in the immediate act of the performance. This frequently lauded “spirituality” of the paintbrush, which actually justifies the equation brush=artist, is nevertheless called into question at the end of the eighteenth century. The admirableness of stretching the rules, that Bettinelli and Cochin took as an sign of genius,30 since this brings out the intemperance and arbitrary nature of the artistic temperament, is questioned by the painter Joshua Reynolds in his posthumously published Discourses on art (1797). The deliberate bending of the rules of art is, according to him, cultivated by contemporary art in order to assert oneself as a genius. Against a spiritualization of the brush stroke, Reynolds emphasizes the instrumental character of the “painter’s language,” which he considers “bare” (whereas according to him only the conviction that this language expresses is to be seen as the work of art). Specifically, Reynolds is highly critical of the Venetian language of painting that is so connected to the thematization of the brushstroke. It is dismissed as “mere struggle without effect, a tale told by an idiot, full of sound and fury, signifying nothing.”31 This undoubtably emotionally loaded basic criticism can be explained as a result of the success of this school that according
28
29
30 31
dente, una nazione più ardita, più mobile, più loquace si dicono predominante dall’Entusiasmo; ed esso viene attribuito al Guerriero principalemente.” Bettinelli. Dell’Entusiasmo. 8. “Il n’y a point de maître plus étonnant que le Tintoretto. L’enthousiasme de son génie & la fureur de son pinceau sont au dessus de toute comparaison.” Christian Michel. Le Voyage d’Italie de Charles-Nicolas Cochin. Rome: École française de Rome, 1991 [1758]. 380. “Ce grand feu fait qu’en général le Tintoretto est moins admirable quand il est fini; son imagination se refroidit pendant le temps de l’exécution; & comme il n’a pas la correction du dessein, & le sçavoir de détail, qui est la perfection de l’exécution finie, il lui reste peu de beautés.” Ibid. 343. “Il étoit tellement plein d’enthousiasme, qu’il n’a pu se contenir dans les bornes de la raison: mais ces écarts sont dignes d’admiration.” Ibid. 342. Sir Joshua Reynolds. Discourses on Art. Ed. Robert W. Wark. San Marino: The Huntington Library, 1959. Discourse IV (1771). 64.
“Il pennello artificioso”
121
to Reynolds had corrupted the younger generation. In a later conference he reflected: “We are charmed with such an unexpected happiness of execution and begin to be tired with the superfluous diligence, which in vain solicits an appetite already satiated.” The allegation of the pure superficiality of Venetian painting – and that means in the narrowest sense ‘meaninglessness’; it is also dismissed as ‘ornamental’ – is clarified by Reynolds’ separation of an “industry of the hand” from an “industry of the mind” whereby the latter guarantees the right to produce art. Rather than finding that it has its own inherent intelligence, the lists of artistic skills are morally dismissed as “tricks of mechanical habits.” As an argument against ‘true’ excellence in painting, he characterizes the facile manipulation of the brush as the “genius of mechanical performance” or “mechanical genius.” It “operates without much assistance of the head.”32 In his first academic lecture Reynolds rebukingly points out that the young artist must be wary of the allure of such ingenious looking brushwork, the imitation of which – fatally according to Reynolds – promises a quick gain in original artistic expression. The glamorous ‘facility’ that the budding painter favors over a painful and humbling diligence in the execution of the work is accordingly discredited by Reynolds as “corrupt” and “frivolous,” since it deliberately conveys the wrong impression. The impatient student wants to “take the fortress by storm,”33 but fails to see that this language is not to be learned “in just a few steps.” The fact that the “facility” of the execution is accorded the highest admiration can be deduced from records of the contemporary criticism of Simon Vouet’s works of art. His reputation as the most influential artist of his times – he was described by Charles Perrault as “le Restaurateur de la Peinture” – is based on the “freedom and freshness of his brushes.”34 De Pils mentions in his Abregé de la vie des peintres (1767) that the
32 33
34
“That kind of mechanick Genius which operates without much assistance of the head.” Ibid. Discourse XII (1784). 216. “The impetuosity of youth is disgusted at the slow approaches of a regular siege, and desires, from mere impatience of labour, to take the citadel by storm. They wish to find some shorter path to excellence, and hope to obtain the reward of eminence by other means, than those which the indispensible rules of art have prescribed.” Ibid. Discourse I (1769). 18. “Quelque habile qu’il ait esté dans son Art, on peut dire cependant que son plus grand mérite consiste dans la grand nombre d’excellens Eleves qu’il a faits, & d’avoir esté en quelque sorte le Restaurateur de la Peinture . . . C’est qu’il avoit la plus recommandable estoit la liberté & la fraicheur du pinceau, qui charmoit le veue par la vive opposition des ombres & des lumières, quoy que pour l’ordinaire elles
122
Nicola Suthor
King so marveled at Vouet’s facilité that it gave him pleasure to watch him working and he also desired to be taught the art of pastel by him. The article dedicated to “facilité” in Watelet’s Dictionnaire is evidence of its positive standing. By way of introduction, the natural gift of the “génie” is called in to explain the “légèrté d’un pinceau facile.” This “fortunate skill” must nevertheless be rebutted at the point at which the material used by the genius is also readily available. Watelet distinguishes between two forms of “facilité”: that of composition and that of painting practice. The “freedom and candour” (“la liberté, la franchise”) of the brush must nevertheless be restrained by the correct use of the medium, otherwise the “génie libertin” impinges on and provokes lapses in the art (significantly the antonym for “facilité“ is “la gêne”).35 “Le faire” – defined as a mechanism of the brush and hand36 – shows itself above all in the “boldness” of the brush (hardiesse), that is defined as an effortless handling of the instrument and is ranked beside “liberté” and “nouveauté” to characterize “genius.”37 But the style of the representation must conform to the idea of the image – the “tout ensemble” – in other words, it must be inspired by this. This kind of “manœuvre intelligente & facile”38 requires the hand to be trained in the handling of the instrument that ultimately defines the uniqueness of the manière of every single painter. Thus the painter must bend (plier) his hand for quick and simple use of the brush.39 This “operation” must be continually repeated to become second nature to him, which is the routine that guarantees a facility in the act of painting. The conditioning of the hand in the use of the instrument cultivates one’s own manière, since it ultimately gives rise to one’s own language, which in turn can be adopted – almost without reflection – by others once it becomes convention. François Dandré Bardon ultimately distinguishes between the very manière that attests to artistic uniqueness and the academic manière
35 36
37 38 39
fussent un peu trop fortes & trop marquées.” Charles Perrault. Les hommes illustres qui ont paru en France pendant ce siècle . . . Paris: Dezallier, 1700. 90. Pierre Charles Lévesque. “Gêne.” Watelet and Lévesque. Dictionnaire des arts. Vol. 2, 39ff. “C’est la pratique de la peinture, c’est au mécanisme de la brosse & de la main, qui tient principalement cette expression.” Dandré Bardon. “Faire.” Watelet and Lévesque. Dictionnaire des arts. Vol. 2, 258. Pierre Charles Lévesque. “Génie.” Ibid. Vol. 2, 396. Bardon. “Faire.” Ibid. Vol. 2, 360. “Pour acquérir la facilité nécessaire à l’exercice des talens, nous sommes obligés de répéter une infinité de fois les mêmes opérations, les mêmes mouvemens, les mêmes procédés, & de plier, par une langue habitude, nos organes à l’emloi que nous voulons leur donner.” Claude-Henri Watelet. “Manière.” Ibid. Vol. 3, 376.
“Il pennello artificioso”
123
of the adepts when examining the difference between “avoir une manière” and “avoir de la manière”, where the latter formulation is unmistakably regarded negatively.40 The ungrounded adaptation of a painterly style that does not develop from one’s own hand but that nevertheless, by the allure of its surface, invites one to “mechnically” trace it, is the basis of Reynolds’ fundamental criticism. Linked to this is the conflict between ars and métier, which towards the end of the eighteenth century and the pre-modern era almost returns art discourse to its beginnings.41 The extraction of a concept of art via the ‘segregation’ of ars from the craft – programmatically by the release of the artist from the guild system – resulted in a lessening of the significance of techné (as not only the etymological root of ars) in art discourse in favor of an appreciation for design – the intellect-driven precursor of the actual painting – as the basis of art.42 The assertion of painting as a liberal art, as Paolo Pino mentions in 1548 in his Dialogo di pittura, is bound up with a reduction in manual execution: “Painting requires no physical labour.”43 After all, the intelletto shies away from this mechanical and laborious aspect of artistic activity.44 When Leon Battista Alberti claims in his 1435 treatise De Pictura: Whether you practice painting or sculpture, you should always have before you some fine and remarkable model which you observe and copy; and in copying it I believe that diligence should be combined with speed of execution, but in such a way that the painter will never apply his brush or style to his work before he has clearly decided in his own mind what he is going to do and how he will do it. It is safer to remove errors with the mind than to erase them from one’s work.45
then it becomes obvious that prestezza does not come from a training of the hand, but rather that the intellect guiding the hand – “quick off the 40 41
42
43
44 45
Bardon. “Faire.” Ibid. Vol. 1, 27. Uwe Fleckner. “‘Pourquoi une belle esquisse nous plaît-elle plus qu’un tableau?’ Fragonard, Diderot et l’éloquence du pinceau dans quelques portraits du XVIIIe siècle.” L'art et les normes sociales au XVIIIe siècle. Ed. Thomas W. Gaehtgens. Paris: Editions de la Maison des sciences de l'homme, 2001. 509-33. “In the case of style, however, it was a truth that definers wanted to avoid, because the ‘mano’ of ‘maniera’ could also be construed as a reference to craft and manual production, associations that artists had fought long and hard to overcome.” Sohm. Style in Art Theory. 74. “E perché la pittura non vuol laboriosità corporale.” Paolo Pino. Dialogo di pittura di Messer Paolo Pino [1548]. Ed. Susanna Falabella. Rome: Lithos editrice, 2000. 132. Concerning this, cf. Martin Warnke. The Court Artist. On the Ancestry of the Modern Artist. Cambridge: Cambridge University Press, 1993. 165. Pino. Dialogo di pittura. 126. Leon Battista Alberti. On painting. Trans. Cecil Grayson. Ed. Martin Kemp. London: Penguin, 1991. 92.
124
Nicola Suthor
mark and not to be held up by any difficulties of execution”– is alone in control of art. Hence, while in the fifteenth and sixteenth centuries the praise of speed proclaims the triumph of painting guided by reason as a liberal art superior to the crafts, defining prestezza in seventeenth century art history is focused on the creative potential that the manual activity genuinely produces. The relationship of mens and manus that marks the various concepts of art can be evaluated by means of the specific scope given to the handling of the instruments. The specialization of the hand by the use of the instrument creates a “scientification” of the practice; painting is ultimately described as “science.” The question of whether imitation can become a science governed by its own principles was already argued by Augustine in terms of the use of the brush. In the first book of De musica he points out the necessity of training the fingers in order to play a musical instrument. By repeating specific exercises the individual participants are trained and the hand is ultimately capable of “imitating” music. The speed and facility that is the goal of this training is nevertheless a result of practice, not of the science that it is based on. But is the imitation really irrational because it can only be explained by the fact that the body is conditioned? Or rather, does this training not generate its own specific knowledge, which can become the basis of a concrete science, since it alone creates rationality? Augustine explains the problematic relationship of mens-manus in art practice by the fact that “imitation does not exist without the body,” but that science “comes from the intellect.”46 Expert knowledge nevertheless develops as a consequence of memory and the senses, both potentials of man that are based in concrete physicality and not in the mind (Augustine also ascribes this to animals). Despite the light and efficient mobility of the finger – it is guided by the intellect, but it is a “thing of the body”– the musician can be inferior in the science. This bodily-saturated knowledge that comes from experience explains not just the ability of musicians and carpenters, but the success of surgeons who also draw from this treasure trove. The word itself – cheirourgein: to perform with the hands – serves Augustine’s illustration of the fact that the scholar possessing knowledge, but not experience, is inferior to the “craftsman” in the implementation of practice-based knowledge. The development of expert knowledge that has to hold its own not in the abstract but in a concrete situation, by the use of the instrument, creates a close bond between body and instrument, that is adapted to the 46
Aurelius Augustinus. De musica. Ed. and trans. Frank Hentschel. Hamburg: Meiner, 2002. Book 1. 136ff.
“Il pennello artificioso”
125
tool and conditions the body. The momentous nature of this relationship is pointed out by Edgar Wind in Art and Anarchy: “Any instrument, it is true, brings with it the danger that is might enslave the man that it is meant to serve. Thus the pencil dominated Meissonier, Menzel, and Muirhead Bone, and their perception became as mechanical as their skill.”47 This “enslavement,” determined by a devotion to the medium of painting as the potential of art, could have been added to both the dominant categories of art discourse of the period, diligenza and prestezza, whose extreme positions were criticized by art theory as mindless. The lack of expert knowledge in the former case, so that the process of painting – becoming independent – never comes to a “fortunate” end, is placed in opposition to the excessive use of the medium. This second excess is not regulated by the intellect either and indulges in the superficial (Reynolds’ argument). The forced enclosing (renfermer) of the artist by practice in his own manière is nevertheless not just self-enslavement, but also the necessary condition of the – more or less theatrical – performance of self as a creative subject. This unawareness of the actual medium being used – the basis of any virtuosity (the instrument must be forgotten) – is not just necessary in order to “speak” for oneself, it is also the condition of that artistic freedom, that in the seventeenth century is regarded as the basis of artistic creation.
WORKS CITED Alberti, Leon Battista. On Painting. Trans. Cecil Grayson. Ed. Martin Kemp: London: Penguin, 1991. Algarotti, Francesco. Saggio sopra la Pittura. Ed. William Spaggiari. Rome: Archivio Guido Izzi, 2000. Augustinus, Aurelius. De musica. Ed. and trans. Frank Hentschel. Hamburg: Meiner, 2002. Baldinucci, Filippo. Vocabolario toscano dell’arte del disegno . . . Florence, 1681. Bardon, François Dandré. Traité de Peinture suivi d’un essai sur la sculpture pour servir d’introduction à une histoire universelle, relative à ces beaux-arts. 2 vols. Paris, 1765. 47
Edgar Wind. “The Mechanization of Art.” Art and Anarchy. 83. Sohm inverts the proportion of conditioning and points out the incorporation of style in the body. He states: “According to Wölfflin, the notion that ‘every painter’ paints ‘with his blood’ has been known and accepted for a long time. The body-style metaphor remains useful today. Barthes says that style ‘springs from the body’ as a kind of ‘reflex’ and biological ‘substance.’ Wollheim denies that style must have a ‘reflective consciousness’ and argues instead that it can be ‘encapsulated in the artist’s body’ as a pattern of manual control.” Sohm. Style in Art Theory. 70.
126
Nicola Suthor
Bettinelli, Saverio. Dell’Entusiasmo delle belle arti. Milan, 1769. Boschini, Marco. La carta del navegar pitoresco. Ed. Anna Pallucchini. Venice: Istituto per la collaborazione culturale, 1966. Dominici, Bernardo di. Vite de’ Pittori, Scultori ed Architetti napoletani. Naples, 1742-45. Fleckner, Uwe. “‘Pourquoi une belle esquisse nous plaît-elle plus qu’un tableau?’ Fragonard, Diderot et l’éloquence du pinceau dans quelques portraits du XVIIIe siècle.” L'art et les normes sociales au XVIIIe siècle. Ed. Thomas W. Gaehtgens. Paris: Editions de la Maison des sciences de l'homme, 2001. 509-33. Kanz, Roland. Die Kunst des Capriccio. Berlin: Deutscher Kunstverlag, 2002. Klingsöhr-Leroy, Cathrin. Das Künstlerbildnis des Grand siècle in Malerei und Graphik. Vom “Noble Peintre” zum “Pictor Doctus.” Munich: Fink, 2002. Lebrun, Pierre. “Recueil des essaies des merveilles de la peinture de Pierre Lebrun, Peintre (Brussels Manuscript 1635).” Medieval and Renaissance Treatises on the Arts of Painting. Original Texts with English Translations. Ed. Mary P. Merrifield. New York: Dover, 1999. 766-841. Lemoine, Annick. “Nicolas Régnier, zugeschrieben, Porträt des Marquese Vincenzo Giustiniani.” Caravaggio. Die Sammlung Giustiniani und die Berliner Gemäldegalerie in Preussen [Exhib. cat.]. Ed. Silvia Danesi Squarzina, Milan: Electa, 2001. 198-201. Michel, Christian. Le Voyage d’Italie de Charles-Nicolas Cochin. Rome: École française de Rome, 1991 [1758]. Orlandi, Pellegrino Antonio. Abecedario pittorico. Venice: Pasquali, 1753. Perrault, Charles. Les hommes illustres qui ont paru en France pendant ce siècle . . . Paris: Dezallier, 1700. Pino, Paolo. Dialogo di pittura di Messer Paolo Pino [1548]. Ed. Susanna Falabella. Rome: Lithos editrice, 2000. Reynolds, Sir Joshua. Discourses on Art. Ed. Robert W. Wark. San Marino: The Huntington Library, 1959. Ridolfi, Carlo. “Vita di Tintoretto.” Le Meraviglie dell’Arte ovvero le vite degli illustri pittori veneti e dello stato. Vol. 2. Padua: Arnaldo Forni Editore, 2002 [1648]. Rossi, Massimiliano. “La peinture guerrière. Artistes et paladins à Venise au XVIIe siècle.” La Jérusalem délivrée du Tasse. Poésie, peinture, musique, ballet. Ed. Giovanni Careri. Paris: Klincksieck, 1999. 67-108. Scannelli, Francesco. Il Microcosmo della Pittura. Bologna: Nuova Alfa Editore, 1989 [facsimile of the edition Cesena, 1657]. Scaramuccia, Luigi. Le finezze de’ pennelli italiani, ammirate, e studiate da Girupeno . . . Pavia, 1674. Sohm, Philip. Style in the Art Theory of Early Modern Italy. Cambridge: Cambridge University Press, 2001. Vasari, Giorgio. Le Vite de’ più eccellenti pittori, scultori e architettori. 6 vols. Ed. Rosanna Bettarini and Paola Barocchi. Florence, 1976. Velasco, Don Antonio Palomino. Las vidas de los pintores y estatuarios eminentes españoles . . . London, 1724. Warnke, Martin. The Court Artist. On the Ancestry of the Modern Artist. Cambridge: Cambridge University Press, 1993. Watelet, Claude-Henri and Pierre Charles Lévesque, eds. Dictionnaire des arts de peinture, sculpture et gravure. 5 vols. Paris: Prault, 1792. Wind, Edgar. Art and Anarchy. Illinois: Northwestern University Press, 1985. Zanetti, Antonio Maria. Della pittura veneziana. Venice: Albrizzi, 1797.
BARBARA MARIA STAFFORD
The Enlightenment “Catholization” of Projective Technology: Theurgy and the Media Origins of Art
Roman epic is painted/ in black fire on black ground . . . Rosanna Warren, Bonfires1 But I do believe, just as Plato, that there are certain divine powers placed between man and the gods which concerned both their nature and their location, and that these govern every divination and magical miracle. (Apuleius, Apology)2
Magic, as Hans Dieter Betz reminds us, is a venerable worldview, one that tries to understand things in their mutual, but ordinarily concealed, connectedness.3 But, today, we have forgotten for just how long optical devices – or tools for “forcing the gods” – have been employed in a wide range of interactive sacred practices. These instrumentalized body-based routines included casting spells, controlling ghosts,4 conjuring the dead, and projecting mortals into the foggy regions of the immortals or bringing the bright gods down to earth. Mirrors, lanterns, diaphanous textiles, and black boxes of all sorts, then, need to be inscribed within the comparative history of aesthetics and religion as well as of science 1
2 3
4
From the collection of poems by Rosanna Warren. Departure. New York: Norton, 2003. See the review by Charles Simic. “Difference in Similarity.” New York Review (March 11, 2004): 21-23. Apuleius. The Apologia and Florida of Apuleius of Madaura. Trans. H.E. Butler. Westport: Greenwood Press, 1970. Chapter 43, 55. Hans Dieter Betz. “Magic and Mystery in the Greek Magical Papyri.” Magika Hiera. Ancient Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 246. Also cf. my Visual Analogy. Consciousness as the Art of Connecting. Cambridge and London: MIT Press, 1999. Sarah Iles Johnston. “Songs for the Ghosts.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 92-94.
128
Barbara Maria Stafford
and technology. Media, magic, and mechanism are inextricably interwoven in an enduring narrative of allure and deviance. I argue that during the Enlightenment this age-old spiritualization of optical technology becomes “catholized” – that is, made part of the larger hostility or discomfort with visual representation found in Protestant post-Reformation Europe. Image-mongering can thus be distinguished from legitimate science or even Jesuit science. While recent scholarship has complicated our ideas of what constituted the “iconic” in early-modern Europe, transformations of visual culture tend to be analyzed from within the delimited fields of history, art history, religion, and culture studies. In reality, tools, ritual practices, and visual or performative environments cannot so easily be pried apart. In contrast, this essay sets out to demonstrate the elaborate intersections of theater, theology, technology, and aesthetics. I propose that projective devices, in particular, become mobilized to tell a graphic story of the international culture of oriental despotism, global priestcraft, and the persistence of illusory styles of superstition. Animation is central to theurgy. This bringing-to-life of the gods in tableaux vivants, both literally and metaphorically, projects a concealed power over illiterate populations. Phantasmagoric apparatus materializes an otherwise immaterial ontology; it corporealizes even the most esoteric philosophical system. Closer to home, the spectral illusions it generates invade the frame of our inner life, perforating the structure of consciousness itself. If technologically produced images – or media – are also mental pictures that exist apart from, behind, or beyond the natural images received through the senses, then vision is equally double. Ignoring physical boundaries, artificially enhanced sight is truly open-ended: snatching a luminous presence from outer space to relocate it within ordinary life or, conversely, chasing after an alternate world towards which we ceaselessly stretch. Founded on geocosmic principles of sympathy and antipathy, the binding rituals of ancient magic5 arose within a mantic culture. Mantics, as opposed to hermeneutics, is fundamentally performative in origin rooted in the articulatory gestures and hand movements of early cultures.6 It requires corporeal expertise in dynamic divination routines and the expansion and extension of the borders of the material body (turning it into a communication technology for hooking up with an5 6
Christopher A. Faraone. “The Agonistic Context of Early Greek Binding Spells.” Magika Hiera. 3-7. On mantics, cf. my essay, “Levelling the New Old Transcendence. Cognitive Coherence in the Age of Beyondness.” New Literary History 35.2 (2004): 321-28.
The Enlightenment “Catholization” of Projective Technology
129
other being or the beyond). Significantly, recent advances in brain imaging, neuroscience, and genetics – probing the first languages – have discovered fascinating links between such seemingly unrelated abilities as mimicry and movement. Michael Arbib, David McNeill, and Steven Pinker are arguing that, although we tend to associate language first with sound rather than movement, it now seems that speech may be better understood as a motor activity.7 I am terming mantics precisely this imitative ability to combine speech and gesture under a meaning – an idea that threads its way from Vico to Rousseau. Magical “self-help” procedures, specifically, interwove manual with oral behavior. This process of reaching-out literally to the alien or the unknown began within “Orphic” or shamanistic societies agonistically ruled by competitive strategies for survival and dominance. The shaman, like his kindred shady cousins, the magician and the priest, fused persuasive saying and doing to his own person, somatically appropriating this special power either for collective or personal ends. Highlighting the lopsided relationship obtaining between the anti-social keeper of secrets and those who would freely know, Plato refers derogatorily to the wandering magician whose spoken formulas and choreographed movements come only at a price.8 From Herodotus to Heraclitus the Ephesian to Xenophon down to Clement of Alexandria, the magos symbolized the embodiment of Persian priestcraft. Belonging to the dubious tribe of “experts” in everything touching the gods, these “specialists” were hardly seen as perfect sages, but rather more as “vagabonds of the night” who arrogantly claimed the divinities were in their service.9 The use of illusionistic devices, instead of words, by manipulative magoi was a key characteristic distinguishing the inferior magic of the juggling prestidigitator or fraudulent wonder-worker (goes) from the superior, learned magician-philosopher of the Hellenistic period.10 Manipulating the gods and their believers in a cunning way thus long prefigured the theurgic contamination of late Neo-Platonism. In the dark writings of Iamblichus, Proclus, and the Pseudo-Dionysius – with their visionary travels to the underworld, mediumistic trances, and mysterious animation of statues – the 7 8 9 10
Constance Holden. “The Origin of Speech.” Science 303 (February 27, 2004): 131619. Plato. Republic. Indianapolis: Hackett, 1992. 364C. Fritz Graf. La magie dans l’antiquité greco-romaine. Idéologie et pratique. Paris: Les Belles Lettres, 1994. 31-33. Matthew W. Dickie. “The Learned Magician and the Collection and Transmission of Magical Lore.” The World of Ancient Magic. 164.
130
Barbara Maria Stafford
“pure” philosophy of Plato appeared to degenerate into a confused mélange of negative philosophy, mysticism, and necromancy.11 Starting where rational reflection ended, a rapturous ritual discipline won out over mere intellect, transporting the enthusiastic viewer into another sphere. By the first century, Pliny, who loathed the empty pretensions of the miracle-mongering Zorastrians, as well as Ireneus, the fulminating Christian Bishop of Lugdunum (active in the latter part of the second century), compiled a considerable catalogue of optical tricks that duped the eye. These praestigiae ranged from how to make an egg look like an apple or how to position the smoking head of a hare so as to make gladiators in a painting look like they were fighting one another. But equally important for the subsequent, virulent Enlightenment critique of priests as mystagogues was Pliny’s globalization of magic. In Book XXX of the Natural History, he lambastes this fraudulentissima artium originating in the invidious combination of quack medicine, ecstatic religious beliefs, and the prognostications of astrology. Born of manipulation, such superstition menaces the measured relationship that normally unites human beings to the gods. Anticipating Marcel Mauss’s categorization of the magician as simultaneously priest/philosopher/charlatan/charismatic, Pliny argues that this secret sect of adepts possessed certain shared features visible from Pythagoras to the Druids. Notably, their impious techniques were foreign imports and their sacrifices relied on divination requiring “professionals” in rites to perform them.12 The Roman author thus internationalizes this epidemic of irrationalism. He vividly impresses upon later readers how the magicas vanitates spread from Persia to Greece, and then infected the Carians, the Cypriots, the Jews, before invading Italy, Gaul, and Britain. It was left to the philosophes, however, to work out the despotic mental and social consequences of believing in miracle workers and their “spiritual machines.” This essay, then, is a chapter in the still unwritten history of the orientalization of optical technology. Enchanting, ensnaring, and drug-like remain apt descriptions not only for its occult baroque, but for its neo-baroque, or twenty-first century, immersive form.13 Virtuality’s fantastic special effects14 depend upon 11 12 13
Jens Braarvig. “Magic. Reconsidering the Grand Dichotomy.” The World of Ancient Magic. 46-49. Graf. La magie dans l’antiquité greco-romaine. 61-75. For an expansion of Omar Calabrese’s concept of the neo-baroque as a system displaying a loss of totality in favor of instability, change, and polydimensionality, cf. Angela Ndalianis. Neo-Baroque Aesthetics and Contemporary Entertainment. Cambridge and London: MIT Press, 2004. 15-19.
The Enlightenment “Catholization” of Projective Technology
131
the seductive persuasion of hermetic experts and the participatory enthusiasm of charmed initiates. It is not surprising that rational critics have always been squeamish about intercourse with the beyond and the marketing of the paranormal. One of the chief characteristics of the Enlightenment was its abhorrence of dark superstition, “Asiatic” idol worship, and the “Oriental despotism” of hoodwinking priests, especially fanatical Roman Catholics.15 Second only to anti-Semitism in power and longevity, anti-Catholicism was embedded in a primal myth of visual and technological iniquity stereotyping shady foreigners: dissolute Italians, wanton French, cruel Spaniards, given to trafficking in the detested manifestations of the otherworldly. During the later seventeenth and eighteenth centuries, this cultural paranoia increasingly demonized the Church whose machinations seemed epitomized in the shifty mirrors, metamorphic lenses, distorted perspectives, and “thaumaturgical projectors,”16 invented or perfected by the Jesuits. I want to argue, however, that, for the Enlighteners, lurking behind the artifices of Catholic spectacle and its excess of fictive representations was a worse visual chaos threatening to spill out. I refer to the entire ominous mantic universe of “enthusiastic” performance with its congeries of simulations.17 Magic lanterns, shadow projections, and nebulous transparencies were invested with some of the same conjuring powers found in theocratic religious ritual which it seemed useless to resist. This machinic expansion of space – identifiable with the virtuoso swell and hyperbole of Baroque aesthetics – violated the controlling orthogonals of linear perspective. The synthetic extension of the finite into the infinite encouraged the secular to commingle promiscuously with the sacred instead of being gridded into discrete and separate domains. As Hubert Damisch proposed, classical perspectiva artificialis imitated vision no more than the Florentine painting of the fifteenth century imitated space. Rather, its epistemological as well as ontological function was to make evident the (proper) graduated procession 14 15 16 17
Janet H. Murray. Hamlet on the Holodeck. The Future of Narrative in Cyberspace. Cambridge and London: MIT Press, 1997. 97. Cf. my Artful Science. Enlightenment Entertainment and the Eclipse of Visual Information. Cambridge and London: MIT Press, 1996. 3-23. Athanasius Kircher. Ars Magna Lucis et Umbrae . . . Amsterdam: Ex Officina Janssonio-Wassbergiana, 1671. 768. On the fascination of myth for the pre-romantic generation stretching from Vico to Paul Henri Mallet, James Macpherson to Johann Georg Hamann, and Johann Gottfried Herder, cf. Bruce Lincoln. Theorizing Myth. Narrative, Ideology, and Scholarship. Chicago and London: University of Chicago Press, 1999. 50-54.
132
Barbara Maria Stafford
of the visible as it becomes apparent to the fixed eye.18 A key benefit of this geometric optic is that we witness how the subject is, in fact, held captive inside the boundaries of the visual field. I suggest that it is precisely this imprisonment and immobilization of the delimited viewer that projective technology, in particular, fundamentally challenged. In Leibnizian fashion, earth-bound beholders became mobilized into distributed and lofty monadic points of view. With the aid of an instrumentally amplified vision, individuals were now able to wander disembodiedly through a plurality of worlds. Conversely, and not unlike those ancient magic acts binding this world to the next – performed by an Indo-Iranian priestly caste and their Persian descendants – projective apparatus deployed an exotic cosmology and injected foreign, grotesque, or demonic forces into ordinary life. Probing and penetrating equipment colonized novel geographies by bringing strange beings inside the borders of the known world. The phantasmagoric shades of the departed were drawn back to earth, fiends could be cast against the wall, and mortals were thrust into the realm of the immortals. Against the backdrop of global magic, Catholicism was merely the most recent instantiation of this animating theurgy of special effects. Identified with the stealthy and enervating automatism of bombarding media, the darkened Wunderkammer Church became the backward antagonist against which a progressive, illuminated, and secular society defined itself.19 For the Enlightenment, it was specifically baroque projective technology that evoked those twin “Romish” excesses of the pagan and the alien. Catholicism’s worldly syncretism, its easy – some would say facile – blending of ancient and modern symbols into a unified sensual ritual, was anathema to an orthodox view of an original, pure Christianity sealed off from heterodox foreign influences. The Church’s analogical integration – and thus facile transformation – of systems of belief that, on the surface, were deeply at odds with one another was unnervingly mimicked by combinatorial instruments that collected, synthesized, and displayed strange sights at a distance. Image manipulation was simply the mechanistic corollary to Catholicism’s apparently anamorphic practice of metamorphosing distinctive doctrinal truths into monstrosities. In addition to such dogma-skewing, nothing seemed more dubious to Protestant critics than projective equipment that appeared to 18 19
Hubert Damisch. L’origine de la perspective. Paris: Flammarion, 1987. 55-57. Raymond D. Tumbleson. Catholicism in the English Protestant Imagination. Nationalism, Religion, and Literature, 1660-1745. Cambridge: Cambridge University Press, 1998. 13-17.
The Enlightenment “Catholization” of Projective Technology
133
make the absent present. Concave mirrors and virtualizing lenses not only falsely conflated nearby with faraway phenomena, but in a damning comparison to the sacrament of the Mass, they materialized the immaterial to enrapture an unwitting audience. It is important to note that a major characteristic distinguishing solitary reflective media, such as small mirrors, from group projective media, such as magic lanterns and silhouette-casting lamps, is the different roles they assigned to participatory witnessing. Projection and screen technologies stimulate multiple spectators to abandon themselves together to dreamlike scenarios as they watch luminous or shadowy visions brought to life by moving across a common interface. The Enlightenment attack against a clergy-managed worship was directed at this wizardry of visualization holding ecstatic sectarians in thrall. The word magic, as we saw, comes from the Persian magoi – those itinerant specialists in the invisible. This secretive tribe of experts on anything pertaining to the gods and their realm also connived to seize a large portion of temporal power.20 Not just Protestants, but dissident Catholics, reproached an exploitive priestcraft trafficking in the marketplace of immersive media. “Wherever the vulgar [in ancient times] were astonished at the effects of shadows, electricity, mirrors, and the magnet, interested persons endeavored by these to frighten them; and thus misapplied the powers of nature to promote their own advantage.” This instrumentalized “art of exhibition” – transporting the beholder into a hyper-illusionistic realm – was condemned as being no better than the fairground mountebank’s sleights of hand performed with cups and balls. Pontiffs and clerics “ought to be detested;” indeed, it would be better for the populace “if they will absolutely pay for being deceived, that they should be exposed to a momentary deception from jugglers than to a continual deception from priests.”21 This pan-European controversy, consisting of thousands of smearing pamphlets,22 did not limit itself to excoriating Rome but denounced theocratic23 fanaticism worldwide. The philosophes, especially, were fasci20 21
22 23
Bruce Lincoln. Priests, Warriors, and Cattle. A Study in the Ecology of Religion. Berkeley, Los Angeles, and London: University of California Press, 1981. 60-61. Johann Beckmann. A History of Inventions and Discoveries. 4 vols. Trans. William Johnston. London: Longman, Hurst, Rees, Orme, and Brown, 1817. Vol. III, 288, 306-309. S.J. Barnett. Idol Temples and Crafty Priests. The Origins of Enlightenment Anticlericalism. New York: St. Martin’s Press, 1999. 8. On theurgy as a development of late Neo-Platonism – laced with the subrational forces of the pagan underworld – cf. Jens Braarvig. “Magic.” 46-49.
134
Barbara Maria Stafford
nated by the theurgic rituals and shadowy mystery cults that threaded their way through the history of early religions. Since the beginning of recorded time apparently, shamanistic hierophants turned large populations into semi-conscious slaves held in thrall to magical operations conducted in murky surroundings. Hume, Voltaire, Boulanger, and Diderot not only wished to puncture the institutionalized trance supposedly holding Papists in its sway, they wanted to demonstrate the longstanding kinship between excessive religious fervor and a tricking illusionistic magic. The larger aim of the Enlighteners, then, was to expose the corruption inherent in all forms of religion that depended on charismatic professionals, including a contaminated Christianity. It, too, was allied with ancient crafty cults relying on optical chicanery and spellbinding miracles performed through hidden techniques. Fragonard’s epic fantasy, the Coresus and Callirhoe, exhibited in the Salon of 1765, brilliantly evokes this cavernous theurgic universe pervaded by the theatrical abandon of insider viewers. Ambiguity reigns. Resembling both a cloudy landscape as well as the smoky sanctuary of Bacchus – half-concealing overwrought devotees – this fitful spectacle suggests an ethereal Dreamtime haunted by an invented cosmology, dark lords, and filmy delusions.24 As we shall see, Fragonard was not alone in invoking the ancient world of tableaux vivants. But he refined the horrific staging of “fatal charades” in Roman amphitheaters where Christians were martyred in the arena in the guise of Dirke, bound to the horns of a bull, or Daedalus, plunging towards hungry jaws with non-functioning wings, or Orpheus devoured by the animals he was supposed to charm – in order to stretch the emotions to their maximum.25 The French artist’s frenetic painting of the seduction of the mind by the hallucinating senses not only highlights the noncritical spectatorship of the actors, but stimulates similar out-of-control transports in the beholder.26 But even more importantly, the daemonic density of the picture’s atmosphere – thick with an ominous haze – summons up memories of a divine hierarchy affirming the special knowl24
25 26
On Fragonard’s painting as an embodiment of Diderot’s style of superstition, cf. my “‘Fantastic’ Images. From Unenlightening to Enlightening Images Meant To Be Seen in the Dark.” Aesthetic Illusion. Theoretical and Historical Approaches. Ed. Frederick Burwick and Walter Pape. Berlin and New York: de Gruyter, 1990. 158-79. Nigel Spivey. Enduring Creation. Art, Pain, and Fortitude. Berkeley, Los Angeles, and London: University of California Press, 2001. 35-36. Eik Kahng. “L’Affaire Greuze and the Sublime of History Painting.” Art Bulletin 86 (March 2004): 108-09.
The Enlightenment “Catholization” of Projective Technology
135
edge of a primordial theocracy whose task it was to keep the endlessly ramifying gods in check. “Theurgy,” or the vile alliance of philosophy with priestcraft (la prêtrise), Diderot asserted, was a “superstition that seeks out darkness (les ténèbres) & retires to subterranean places.” The priestly keepers of this necromantic underground magically summoned hybrid gods to appear, offered obscene sacrifices, and stunned cult members into submission through the use of gaudy devices. Artificially smoothed simulations existed only within the matrix of this phantasmagoric theater walled off from outside verification. The dimness of such stage-like grottoes helped to dupe the viewer who mistook representations for real things. “We see an entire scene passing through our imagination, as if it were external to us, because as long as the illusion lasts, all real beings are annihilated and our ideas take their place: it is only our ideas that we perceive, nonetheless our hands touch bodies, our eyes see animate beings, our ears hear voices.” Like lying poets who collect a grab bag of scattered perceptions and weld them into a larger fiction, “enthusiastic” spectators amplify the erroneous phantasms generated by the wayward senses, making them “gigantic, unbelievable, enormous.”27 Doctor Johnson likewise condemned the projective desire to communicate with what lies beyond commonsense in ill-lit spaces. This fanatical impulse to be transported, to leave behind actual perceptions and enter a seamlessly constructed world just by looking at a cobbledtogether image, “dulls” mental acuity. Even worse, having visual sensations in the absence of solid objects leads to ecstasy, madness, or “enthusiasm,” that “excessive elevation and distraction of the mind.”28 In sum, enlightened thinkers wanted to ferret out the tricking simulations, feverish rites, and misleading wonders – i.e., the implicit Catholicism underlying all cults, old or new, that performed their mystifications in 27
28
Denis Diderot. “Eclecticisme.” Encyclopédie, ou Dictionnaire Raisonné des Sciences, des Arts et des Métiers. 35 vols. Ed. idem and Jean LeRond d’Alembert. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 1751-1780. 271, 283. On the negative connotations of enthusiasm, cf. Shaun Irlam. Elations. The Poetics of Enthusiasm in Eighteenth-Century Britain. Stanford: Stanford University Press, 1999. 23-24. Samuel Johnson. A Dictionary of the English Language. London: Knapton, 1755. 383. For English concepts of religious and aesthetic “enthusiasm” and their connection to seventeenth-century fanatical sects touched by “the shadow of Rome,” cf. Bernd W. Krysmanski. Hogarth’s “Enthusiasm Delineated,” Nachahmung als Kritik am Kennertum. Eine Werkanalyse. Zugleich ein Einblick in das satirischaufgeklärte Denkens eines “Künstlerrebellen” im englischen 18. Jahrhundert. Hildesheim: Olms, 1996. Vol. 1, 688-93.
136
Barbara Maria Stafford
semi-obscurity, thus making contrived objects difficult to see through or see accurately. What makes the specifically baroque, hyper-realistic mapping of mass and volume in perspectival space so satanic to skeptical critics, I believe, is precisely the projective or eruptive property of constructed depth. By projection, I mean the uncanny ability of black-box or cavelike environments to persuade us that the glinting things we see flitting on the stage or the screen or behind the sheet are going to come out. The modern psychology of perception speaks of the powerful “kinetic depth”29 or “structure-from-motion”30 effect whereby movement influences our perception of shape and creates three-dimensional forms on two-dimensional surfaces from moving shadows. Special effects that exaggerate structural motion can give a heightened sense of dimensionality, virtually transporting the viewer into the scene. Extreme conditions of brilliance and obscurity contribute to this kinetic sensation of being involuntarily pulled into an infinite expanse. Religious epiphany appears to be just such a will-less spatio-temporal breakthrough from our circumscribed bodily world into the boundlessness of the otherworldly. The modern, “Enlightened” association of technology with secularization has tended to overlook its historical role in materializing the sacred. Actual projection, or simply feigning the realistic effect of something jutting from one realm into another, was a venerable component of priest-determined worship since it seemed to manifest the gods in their actual presence. A centuries-old tradition linking religion, theater, and magic lantern was based precisely on their combined ability to draw the viewer into an astonishing twilight space susceptible to intense imaging. Like the misty form, or eidolon, sent in dream magic, the incense-burning lamp, or nocturnal sun, generates smoky likenesses hovering “above the head.”31 More generally, the illusionistic tendency to exhibit “living” pictures in the dark – the lambent phasmata of gauzy deities and numinous spirits – provocatively transgressed the line separating art from reality. Equipment artificially exploited the night, with its dead spirits and eerie ways, just as it illuminated the moonless kingdom of the underworld. 29 30
31
E.B. Goldstein. Sensation and Perception. Pacific Grove: Brooks/Cole, 1999. 292. D.C. Bradley, G.C. Chang, and R.A. Andersen. “Encoding of Three-Dimensional Structure-from-Motion by Primate Area MT Neurons.” Nature 392 (April 16, 1998): 714-17. Samson Eitrem. “Dreams and Divination in Magical Ritual.” Magika Hiera. 176-81.
The Enlightenment “Catholization” of Projective Technology
137
The ritualistic and incarnational power of the dumbshow harks back to the simulated stage action of early Egyptian and Greek temple magic. But it is also reminiscent of the mythic spectacula or “fatal charades” of the Roman arenas that tantalizingly coincide with the subjects of certain famous Roman sculpture groups like the Laocoon or the Farnese Bull.32 The memory of these ancient tableaux vivants were alive when the future Lady Hamilton, dressed in a white tunic with a belt around her waist, struck statuesque “attitudes” and underwent refined transformations beneath the shadow of Vesuvius before a host of international notables. Typically, this pageant of animated personifications has been described as Susan Sontag does most recently: “Maids would bring an urn, a scent box, a goblet, a lyre, a tambourine, and a dagger. With these few properties, she took her position in the middle of the darkened drawing room. When the Cavaliere came forward, holding a taper, the performance had begun.”33 Generally unremarked and undeveloped,34 however, is a startling discovery Goethe made during his second Italian trip. It is perhaps not surprising that this reformed illusionist, who now deemed the theater as “merely a peepshow on a larger scale,”35 would become intrigued by the fact that Emma Hart’s kinetic monodrama of Greekish poses had initially taken place in an upright black box. This English girl of twenty was directly illuminated by tall candles that remained concealed during the performance. Inside its shadowy recesses Emma Hart let down her auburn hair, deployed shawls, and silently underwent multitudinous transformations in the auratic enactment of mostly ancient works of art. In a letter to Herder, dated May 27, 1787, we learn that Sir William Hamilton – at the prompting of the German painter Philipp Hackert – showed his secret treasure vault to a small group of guests. This jumbled Wunderkammer was crammed with oddments from every conceivable period: busts, torsos, vases, bronzes, Sicilian agate carvings, paintings, and chance bargains of every sort lay strewn about. There was even a small chapel. In the midst of this confusion, Goethe suddenly glimpses a chest, a sort of decommissioned camera obscura or perhaps a repurposed, walk-in Engelbrecht Theater. 32 33 34 35
Spivey. Enduring Creation. 252. Susan Sontag. The Volcano Lover. A Romance. New York: Farrar, Straus, Giroux, 1992. 145. Ulrike Ittershagen is an exception in noting Goethe’s remark. Cf. her Lady Hamiltons Attitüden. Mainz: Philipp von Zabern, 1985. 46-51. Johann Wolfgang von Goethe. Italian Journey (1786-1788). Trans. W.H. Auden and Elizabeth Mayer. San Francisco: North Point Press, 1982. 187.
138
Barbara Maria Stafford
Its front had been taken off, the interior painted black and the whole set inside a splendid gilt frame. It was large enough to hold a standing human figure, and that, we were told, was exactly what it was meant for. Not content with seeing his image of beauty as a moving statue this friend of art and girlhood wished also to enjoy her as an inimitable painting, and so, standing against this black background in dresses of various colors, she had sometimes imitated the antique paintings of Pompeii or even more recent masterpieces.
This phase, he regretfully concludes, seems over now, “because it was difficult to transport the apparatus and light it properly, and so we were not share in this spectacle.”36 At first reading, this account appears to flatten and restrain Emma’s procession of poses, confining them to a cubicle and thus reasserting their aesthetic distance from the viewer. But, then, almost as an afterthought comes a fascinating aside confirming Goethe’s understanding of these sensible presentations as projective phantasmagoria, that is, as illusionistic tableaux vivants abruptly thrusting an individual outward, life-like, towards a marveling audience. His report also offers an implicit comparison between pagan and Catholic religious practices. He tells Herder that this episode of the chest suddenly reminded him [recall that the black box was discovered in the vicinity of “a small chapel”] of something he had forgotten to report: the Neapolitan’s love of crèches or presepe, visible everywhere in the city’s churches. These constructions consist of groups of large, sumptuous figures representing the adoration of the shepherds, angels, and the three magi. A light framework, like a hut, is decorated with trees and evergreen shrubs. In it, the Mother of God, the Infant and the others stand or float, dressed up most gorgeously . . . In depicting this sacred scene, it is possible that living figures were sometimes substituted for the dolls, and that in time, this gave rise to one of the great diversions of noble and wealthy families, who pass many evenings in their palaces representing profane scenes from history or poetry.37
Emma’s freeze-frame modeling of a Sibyl, Fury, Niobe, Agrippina, Bacchante, Cassandra, Cleopatra, or Isis inside a cave-like device was inspired not only by the Pompeian and Herculean frescoes Sir William Hamilton had engraved. Goethe’s reference to contemporary hyper-illusionistic shrines permits us to situate them within much older oracular religious practices, specifically the fervent, widespread search for divine experience occurring in the Gnostic and Hermetic cults of late paganism that promised intimate communion with the divine world.38 As 36 37 38
Ibid. 310-11. Ibid. 311. Fritz Graf. “Magic and Divinations.” 282.
The Enlightenment “Catholization” of Projective Technology
139
important to the enactment of these lamp charades, I argue, were the mysterious religious allegories of life and death – believed to stem from the early Eleusinian or Cabirean period documented by Herodotus – depicted on the Cavaliere’s famous vases. It is worth exploring an unremarked religious source of such dumbshows and, indeed, of projective apparatus in general. Eleusis occupies a special place in the Western imagination. It stands at the head of all later technologies intent on producing a simulacrum of otherworldliness. The Eleusinian Mysteries were an annual celebration connected to corn festivals, beginning around 1400 B.C. and ending in 396 A.D., when the Goth Alaric plundered and burned the sanctuary. They were performed on the soil of the deme of Colonus, near the rocky hill thought to be the somber entrance to the netherworld.39 The question whether these and perhaps other mysteries received an initial impetus outside of Greece (perhaps Asia Minor, Crete, or Egypt) has still not been clarified. But it is safe to assume that magic was a constitutive component of the rites of mysterium, or the revelation of transcendental realities,40 as Enlightenment authors correctly deduced. Any Greek citizen could be initiated into this primal fertility cult over which Demeter, Persephone (Kore), and her son Brimos presided, but only once in his or her lifetime. Of the experience the initiate was permitted to say he had beheld ta hiera, “the holy.” To say more meant incurring the death penalty.41 The well-known myth of Demeter and Persephone, it has been argued, was the most important myth of classical antiquity to focus on the lives of women. Henri Jeanmaire and others have speculated it was tied to women’s rites. But this hypothesis cannot easily be reconciled with the Eleusinian Mysteries – the most famous ritual associated with the myth – since men as well as women were initiated.42 My interest lies with the scenario. The Greater Mystery was celebrated in September when new candidates would walk the Sacred Road between Athens and the Rarian Plain, finishing the journey by torchlight. Upon arrival, the mystai – dressed in black in the earliest periods – danced late into the night. Only those who had fasted nine days 39 40 41
42
Carl Kerenyi. Eleusis and Athens. Archetypal Image of Mother and Daughter (= Bollingen Series, LXV). Trans. Ralph Manheim. New York: Pantheon, 1967. 85. Betz. “Magic and Mystery in the Greek Magical Papyri.” 250-51. R. Gordon Wasson, Albert Hofmann, and Carl A.P. Ruck. Road to Eleusis. Unveiling the Secret of the Mysteries. New York and London: Harcourt Brace Jovanovich, 1978. 36-38. Bruce Lincoln. Emerging from the Chrysalis. Studies in Rituals of Women’s Initiations. Cambridge: Harvard University Press, 1981. 72-75.
140
Barbara Maria Stafford
could enter the labyrinthine precincts. Once inside, the pilgrims drank the hallucinogenic potion or kykeon – a mixture of mint and ergot-infested barley – passed around by the officiating hierophant. From the Homeric Hymn to Demeter, an anonymous seventh-century B.C. poem, we know this drink induced ventriloquism, rapture, and soul-shattering visions.43 Strong light, bronze gongs, and the pungent scent of frankincense and myrtle floating in the thick air further contributed to the heady atmosphere.44 When we first see Persephone after her springtime abduction, she is confined to the murky gloom of the subterranean realms. Snatched away from light by Hades, the bright sky is taken away from her: an experience of death and rebirth enacted allegorically through initiation. Just as Persephone was snatched away while gathering Gaia’s “flower of deception,” the hundred-headed narkissos, the neophyte has to lose his senses or “die” before being resurrected into a new life. Especially significant for Enlightenment mythographers, concerned with the progress from archaic despotism to civilized society, was the way the barbarous rape and eventual salvation of a young woman came to symbolize a cosmic event: unleashing famine, chaos, overturning the seasons until Persephone/Kore was gradually restored to Demeter (that is, to public life in the city-state) by Zeus. Akin to Lady Hamilton’s living pictures in the dark – whereby she shed her material identity to partake of the lambent experience of a procession of goddesses – all the participants stepped outside their own historical time and place and reentered the numinous time of myth. Conversely, as in the oral-visual culture of early Greece,45 Emma’s isolated and dazzling apparition inside the black shadow box bypassed narrative to be projected directly into the spectator’s present. The culmination of these nine mystical nights was a revelation. Something, apparently, was shown to the crowd assembled outside the telestrion, or somber hall – a cross between a dim temple and a cavernous theater – set into the live rock of the mountainside. Euripides claimed it was the blazing epiphany of Persephone herself. Tertullian dyspeptically claimed it was a phallus; Hippolytus, a grain of wheat.46 In the 43 44 45 46
R. Gordon Wasson. “The Divine Mushroom of Immortality.” The Flesh of the Gods. The Ritual Use of Hallucinogens. Ed. P.T. Furst. New York: Praeger, 1972. 185-200. Maureen B. Cavanaugh. Eleusis and Athens. Documents in Finance, Religion, and Politics in the Fifth Century B.C. Atlanta: Scholars Press, 1996. 191. Jocelyn Penny Small. “Time in Space. Narrative in Classical Art.” Art Bulletin 81 (December 1999): 563. Lincoln. Emerging from the Chrysalis. 86-87.
The Enlightenment “Catholization” of Projective Technology
141
very beginning, it seems, an objectless incandescent “screen” of intense light, fire, and smoke was witnessed outside, at a distance, by those pilgrims who were not authorized to enter the hall. But these empty flashes against the night sky were not yet the ineffable holy things that manifested themselves inside: a brilliant vision of the beginning and end of life melted into a concentrated show of splendor. Whatever else they were, such revelations were simulations, tracing their origins to the illusory stage action of temple magic. They are a vivid reminder of the power once wielded over the observer by dead or obsolete media. Because of the secrecy shrouding the Eleusinian Mysteries and their long and convoluted history, it is unclear whether any devices originally intervened in the generation of visions. In the beginning, did they exist only in the mind’s eye of the narcotic-user who projected sights onto a non-anthropomorphic shining? Plato compares the beholding of the radiant Ideas by the disembodied spirits to the blinding brightness of the mysteries when all veils fall away.47 But by the Graeco-Roman period, there is no doubt that this transcendent flickering had solidified into material images mechanically displayed to the public by priests. There is compelling evidence for the use of sophisticated lighting, gadget wizardry, gauzy veils, trompe l’œil sets. From Herodotus’s and Plutarch’s essays in the first century and, especially, from Apuleius’s picaresque romance in the second century, we glimpse the hold of these false miracles – with their multiple transformations – performed by magicians. St. Augustine disparagingly refers to them as “poetic fictions and theatrical entertainments”48 beguiling the populace. The protean shadow-shows booming in the wake of orientalizing sects49 were filled with shifting shapes and demonic phantasms that defied any form control. As chief priest of Carthage and an initiate into several mystery cults – including that of Isis – Apuleius was probably more knowledgeable about such closed rites than most. Certainly, it was he who popularized this necromantic thaumatury for Roman audiences. In fact, he was brought to trial (although acquitted) for being a magus, not a philosopher.50 47 48
49
50
Plato. Phaedrus. Indianapolis: Hackett, 1995. 249C. Augustine. The City of God against the Pagans. 7 vols. Trans. David S. Wiesen. Cambridge: Harvard University Press, 1968. Vol. III, Book VIII, Chapter XXI, 99. Cf. Luther H. Martin. Hellenistic Religions. An Introduction. New York: Oxford University Press, 1987. 4-10; Richard Fletcher. The Conversion. From Paganism to Christianity. New York: Henry Holt and Company, 1997. 253. Braarvig. “Magic.” 40-42.
142
Barbara Maria Stafford
In his tongue-in-cheek novel, The Golden Ass, an obsessive curiosity (curiositas improspera) is blamed for fueling the mania for gazing at supernatural apparitions.51 This improbable “Milesian tale” cobbles together amusing hearsay, idle stories, and a concoction of old wives’ gossip or aniles fabulae52 – concerning the mock-heroic tribulations of a hapless young man who is accidentally changed into an ass while watching a magic show. After seeing a woman metamorphose herself into an owl with the aid of a miraculous salve, Lucius tries to follow suit but mixes up the ointments to calamitous effect. In order to regain his human form, he must eat rose petals that are maddeningly snatched away whenever he is about to ingest them. This quest for transformation leads to numerous, and often bawdy, adventures. Our interest lies with the tableaux vivants of the final chapter when, on the verge of despair, he is redeemed by Isis into whose cult he ultimately becomes initiated. The procession Lucius attends on the day of his salvation includes not only a parade of costumed people but of concrete symbols. Women dressed in white carry metallic mirrors and brush the empty air with ivory combs in a charade of grooming the goddess. Others sprinkle the streets with perfumes or carry lamps, torches, and candles to honor the queen of the night. Upon their heels follow the material avatars of the gods: Isis in the form of a cow, an urn covered in hieroglyphics, a chest containing mystical emblems. Once inside the dim temple, these “living images” were no longer part of a public display but became incorporated into a private epiphany. This festive spectacle only whetted Lucius’s ardor to be admitted to the solemn mysteries. When the priest Mithras declared the moment to be auspicious, the eager votary was attired in the “Olympic” twelve robes of consecration. “Wherever you looked I was decorated all over with pictures of multicolored animals: here Indian serpents, there Hyperborean griffins with bird-like wings, creatures of another world . . . In my right hand I held a flaming torch and my head was encircled with a beautiful crown of palm, its bright leaves projecting like rays.” Clad in the image of the sun, “I came to the boundary of death and after treading Proserpine’s threshold returned having traversed all the elements; at midnight I saw the sun shining with brilliant light; I approached the 51 52
P.G. Walsh. The Roman Novel. The “Satyricon” of Petronius and the “Metamorphoses” of Apuleius. Cambridge: Cambridge University Press, 1970. 176-77. On the decadent fabulae milesiae of The Golden Ass and its connection to pointless old wives’ tales, cf. Eileen Reeves. “Old Wives’ Tales and the New World System. Gilbert, Galileo, and Kepler.” Configurations 7 (Fall 1999): 304-07.
The Enlightenment “Catholization” of Projective Technology
143
gods below and the gods above face to face and worshipped them in their actual presence”53 (italics mine). Here we discern Apuleius the theurgic Neo-Platonist at work. The journey to enlightenment (photismos), he describes, was arduous: sacred dances and dramas marked the initiand’s travels through the shadowy compartments of the universe. Not unlike later Freemasonry – surely inspired by such colorful accounts of processions with emblems54 – enlightenment was artificially delayed by mystifying occurrences or obstacles to encourage patience and fortitude in the neophyte.55 The seeker was constantly driven onward in pursuit of a higher truth. Nonetheless, without a mystagogue or torch-bearing guide, a password (symbolon), a key, it was impossible to attain revelation. Once having been reborn, i.e., by having traversed the ominous twists and turns of this cosmological dark chamber, the candidate joined the sacred band of “knowers.” Apprehension of the divine secrets, or gnosis, manifested itself in the form of psuche.56 Surely to conjure up these “doubling” powers from the beyond required technological devices. We are told they were like cloudy dream images (oneiros), foggy phantasms, and monochromatic shades (skias). Like the apparatus-augmented tableau vivant, supernatural apparitions were not insubstantial illusions but existed simultaneously on two separate planes: being both opaquely of, and dimly not of, this world. Significantly, Enlightenment critics steeped themselves in such syncretistic late antique sources. These writings were of intense eighteenthcentury interest for several reasons: on the positive side, they created an interface between warring systems of religious belief as well as providing a venerable ancestry for Freemasonry. This vast “philosophical” organization had an elaborate ritual and a moral code of conduct that supposedly went back to Isis and the Cabiri – originally demigods, the sons of the first craftsman, Hephaestus, who shared in the chthonoic rites of Demeter and Kore.57 On the negative side, pagan authors also ex53 54
55 56 57
Apuleius. The Golden Ass or Metamorphoses. Trans. E.J. Kenney. London: Penguin Books, 1998. 199-202, 209. For an account of marches made in the 1770s by the brethren of the Craft carrying emblems, symbols, and banners, cf. John Hamill. The Craft. A History of English Freemasonry. Wellingsborough, Northamptonshire: Crucible, 1986. 77. James Stevens Curl. The Art and Architecture of Freemasonry. An Introductory Study. London: B.T. Batsford, 1991. 34-36. Jean-Pierre Vernant. Myth and Thought among the Greeks. London, Boston, Melbourne, and Henley: Routledge & Kegan Paul, 1983. 308-09. R.E. Witt. Isis in the Graeco-Roman World. London: Thames and Hudson, 1971. 154.
144
Barbara Maria Stafford
posed the devastating connections between priestcraft, optical technology, and polytheism. From the writings of Cicero, Apuleius, and the theurgical Neo-Platonists Porphyry and Proclus – those philosophers of the magic-imbued third and fourth centuries – the philosophes learned that the geography of hell had become populated with Eastern or “Asiatic” gods: Sabazios, Sagreus, Dionysius, Cybele and Attis, Isis and Osiris. To the Baron de Sainte-Croix, Silvestre de Sacy, Sergei Ouvaroff, and James Christie, it seemed that all religions “have had a glimpse of this great truth, the fall of man: it is found in the theological systems of the globe and serves as a foundation for ancient philosophical and mythological traditions.” When such primitive mysteries are artificially personified in tableaux vivants, however, a system of theurgy arises “as in the writings of the principal eclectics” or the “new Platonists.” The immaterial doctrine of Plato, they believed, had become corrupted by “Oriental ideas,” including “the worship of light, the system of emanations, and the doctrine of metempsychosis.”58 Worse, these eclectic Platonists staged materializing rituals. Terrible specters with the face of a dog and other grotesque hybrids stocked their subterranean cosmos where “light and darkness succeeded one another, one could scarcely make out the multiplicity of objects.”59 It has been unremarked in current scholarship about the eighteenth century that the monochromatic Greek vases of the sixth and fifth centuries B.C.E., and their Greco-Roman progeny, were thought to offer rare glimpses into the “Isaic” mysteries at Eleusis. Sir William Hamilton (1730-1803) – envoy extraordinary of His Britannic Majesty to the Kingdom of the Two Sicilies, vulcanologist, lover of the tableau vivantcreating Emma Hart, enthusiastic supporter of the excavations at Pompeill and Herculaneum – was a passionate collector. The famous “Etruscan” vases, assembled from the beginning of his tenure in Naples in 1764, demonstrated to contemporary connoisseurs the tangible “elegance,” “law,” and “rules” governing ancient art. Equally, if not more important, however, was their concrete manifestation of the physical “progress” of the human mind from gullible animism to rational allegory – a tale compellingly told by Hamilton’s French adventurer-scholar, the Baron d’Hancarville.60 58 59
60
James Christie. An Essay on That Earliest Species of Idolatry. The Worship of the Elements. Norwich: Stanhopian Press, 1814. ii. M. Ouvaroff [Sergei Semenovich, Count Uvarov (1786-1855)]. Essay on the Mysteries of Eleusis. Trans. J.D. Price with observations by James Christie. London: Rodwell and Martin, 1817. 31, 65-66. F.A. David. Antiquités etrusques, grecques et romaines. Gravees par . . . avec les
The Enlightenment “Catholization” of Projective Technology
145
Matter always matters to Enlightenment observers. A sophisticated sensitivity to the nuanced empirical properties of materials runs through the commentaries about the antiquarian, natural historical, and industrial objects with which Hamilton, and other dilettanti, surrounded themselves. Even though Naples was a major center for ceramic manufacture during the eighteenth century, it is nevertheless striking how many of the antiquities amassed by him were actually translucent. These included semi-precious intaglio gems and cameos,61 rock crystal, murrhine ware, specimens of polished minerals and lava62 (the so-called “gems” of Vesuvius), and, most spectacularly, the glass (but widely believed to be of onyx or agate)63 Portland Vase. Moreover, like Caylus, d’Hancarville, Winckelmann, and Payne Knight, he was fascinated by how early Christianity appropriated paganism’s material emblems and converted ancient shrines, like the Temple of Osiris at Isernia and elsewhere, to its own immaterial ends.64 Hamilton was not alone in intertwining antiquarianism with esotericism, material science with symbolic stones. His Greekish pots inspired a performative theory of their decoration in which the technical procedure painting acted out aspects of the mysteries that were represented. The flat black or red areas of color, circumscribed by a tight contour, were interpreted as capturing both the iconography and the enactment of the Isaic ritual. “The priests spared neither ingenuity nor expense to establish shews, which, by their splendour and secrecy could charm as well as invite the curiosity of their countrymen,” giving rise “to that elegant class of ancient vessels.” In his extensive study of the (Isaic) Eleusinian Mysteries, James Christie the Younger (1773-1831) declared that Pliny was wrong in thinking the shadows of mortals were original-
61 62 63 64
explications par [Baron] d’Hancarville. Paris: Chez l’Auteur, 1785-1787. Vol. I, 28, 71-72; Vol. II, 7-8. For the connections between Hamilton, d’Hancarville, and fifteenth and sixteenth-century writers on magna Graecia, cf. Claire Lyons. “The Neapolitan Context of Hamilton’s Antiquities Collection.” Journal of the History of Collections 9 (No. 2, 1997): 229-40. On Hamilton, d’Hancarville (Pierre-Francois Hugues), and the publication of the first vase collection, cf. Ian Jenkins. “‘Contemporary Minds.’ Sir William Hamilton’s Affair with Antiquity.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. 45-51, 99-100. Ian Jenkins. “Talking Stones. Hamilton’s Collection of Carved Gems.” Vases & Volcanoes. 93, but the theoretical importance of translucency is not mentioned. John Thackray. “The Modern Pliny. Hamilton and Vesuvius.” Vases and Volcanoes. 67-68. Michael Vickers. “Hamilton, Geology, Stone Vases, and Taste.” Journal of the History of Collections 9 (No. 2, 1997): 270. Giancarlo Carabelli. In the Image of Priapus. London: Duckworth, 1996. 3-5.
146
Barbara Maria Stafford
ly transferred to vases from the example of the silhouette-casting lamp of Dibutades, the Corinthian potter’s daughter. He argued, instead, that the Eleusinian “scenes” were the real prototype for this manufacture of ghostly doubles, proposing that these sacred tableaux were, in fact, “transparencies.” The religious plays took the form either “of a dark superficies, in which transparent figures were placed – hence the Etruscan vases with red figures upon a black background; or of opake figures, moved behind a transparent canvas [a sort of ombres chinoises], and hence – those earlier vases in black figures upon a red ground.”65 Significantly, this gestural theory locates the origin of art not in passive or reproductive copying, but in the performative technology of projective media and within theurgic, or animating, ritual. Eleusis burst into the modern consciousness again in 1814, when English archaeologists conducted the first excavations of its venerable ruins. With renewed intensity, James Christie – the son of the founder of the legendary auction house and an expert on ancient Greek and Italian vases and sculpture – recounted the now venerable story of how conjuror priests transformed the semi-obscurity of the temple interior into a microcosmic dark chamber for “exhibitions with luminous phantoms.” The reader – accustomed to popular magic lantern phantasmagoria, top and back-lit vues d’optiques, and transparencies of every kind – is enjoined to imagine how “the Mystae, introduced in the dark, have taken their seats and wait with trembling expectation for the opening of the mysteries.” Invoking Plutarch’s essay on “Isis and Osiris,” he described how during ceremonies commemorating death and resurrection, statues of the Egyptian goddess were arrayed in partly black and partly white veils. The simultaneous inkiness and radiance of the sacred garment visually captured the essence of the deity. “The mere display of this [dualistic garment] to the people by torchlight, would have produced the effect of a transparency.”66 Importantly, Christie thought that the succession of diaphanous images at Eleusis evoked the reanimation of the universe. Like Diderot, or the Baron d’Hancarville, or Warburton before him, Christie argued that the history of mankind immediately after the universal Flood was recorded in graphic symbols. These early anthropologists believed that mystery cults, ritual artifacts, funerary monuments, 65
66
James Christie. A Disquistion upon Etruscan Vases. Displaying their Probable Connection with the Shows at Eleusis, and the Chinese Feast of Lanterns, with Explanations of a few of the Principal Allegories Depicted upon them. London: T. Becket. 1806. 15, 23f. Ibid. 27.
The Enlightenment “Catholization” of Projective Technology
147
and hieroglyphics – unlike later alphabetical writing – preserved the collective, if confused, memory of humanity’s near extinction and miraculous rescue from the darkness of the deluge. To commemorate this great escape from engulfing chaos, solstitial rites were invented throughout the world to celebrate the wondrous annual return and ascent of the sun from the depths of its winter quarters. Put in terms of black and white, the rising and setting of the sun and moon and, especially, their traumatic eclipses, literally meant that one luminous heavenly body was swallowed or obscured by another. The artifex opifex, or hierophant at Eleusis, “who set the mystic fantoccini [note the disparaging term] in motion and who explained their allegorical meaning”67 resembled those Babylonian astrologers who climbed their seven-storied tower to ponder planetary conjunctions or those Chinese mandarins who ascended their celestial observatory in Beijing, “begun shortly after the Flood,” to check on impending lunar or solar eclipses.68 The Encyclopédie similarly reported that in ancient Thessaly, Egypt, Mexico, and India, shamans used enchantments to lure the moon down to earth during an eclipse to “save” it from being devoured by a dragon-shaped sun and encouraged people to raise a din to drive the attacker away. The Chinese and Romans ignited bonfires and lifted torches to the sky in a desperate attempt to rekindle the extinguished orb.69 Like Goethe’s dyspeptic report to Herder concerning the simulation of sacred stories in modern Neapolitan presepe, these animating charades illuminate a primitive moment in the history of human thought still visible in popular culture. Thus Christie compares the Eleusinian priests who conjure up duping phantasms [fantoccini] to latter-day nomadic mountebanks – continuing to manipulate an ignorant populace as if they were fairground puppets on a string. Less elliptically than Goethe, he implies that we can still deduce from contemporary tragedies and comedies – staged for the marionette theater by perambulating Italian operanti and recitanti – how, long ago, marvelous nocturnal phenomena might have been similarly woven into cosmological dumbshows. These animations brought to life the sun’s or moon’s miraculous rescue from destruction for gullible viewers. 67 68
69
Christie. An Essay on That Earliest Species of Idolatry. 9-11; also cf. his Etruscan Vases. 40. J.B. Du Halde. Description géographique, historique, chronique, politique, et physique de l’empire de la Chine et de la Tartarie chinoise. Paris: P.G. Lemercier, Imprimeur-Libraire, 1735. Vol. III, 339. Diderot and d’Alembert. “Eclipse.” Encyclopédie. Vol. V, 294.
148
Barbara Maria Stafford
The ethical dilemma vexing the philosophes was that the terrifying natural elements responsible for the general catastrophe wiping out humanity according to myth and legend, subsequently became personified as evil demons or good benefactors, creating a living tableau of destroying or preserving beings within an institutionalized creed. Their researches into comparative religious practices – like those of Apuleius earlier – convinced them that there were two kinds of optical technology as well as two kinds of science that the public needed to distinguish. The first type was manipulative and superstitious, arising from the hybrid intersection of religious with political power. It was identified with the supposed special expertise of the magus or priest who claimed to have commerce with the gods but actually “operated” in the dark, behind the scenes, using tricking mechanisms. I have been arguing that the Enlighteners identified this mystifying strand (what I have termed a technological catholicism) with a much older sacred tradition of theocratic kingship and theurgic temple magic. The second type – used to unmask the visual tricks and image-mongering animations of the first – was enlightened, secular, collaborative, involving real knowledge, not illusory gnosis.70 Importantly, these dual conceptions of wonder-working sensory media, while often in conflict, continuously reshaped one another. This coexistence of competing, technically inflected visual environments and rituals during the eighteenth century points up the unsuspected existence of a mystical Enlightenment at the very heart of its rational secular counterpart. Further, the “catholization” of optical instruments as devices – in the double sense of duping apparatus and specious expertise – is still with us today. Perhaps it is nowhere more evident than with the new digital magoi who alone can go below or above the interface to achieve the experience of “real presence” with the gods, i.e., be in direct contact with code.
70
Nicole Fick-Michel. Art et mystique dans les Métamorphoses d’Apulée. Paris: Diffusion Les Belles-Lettres, 1991. 212.
The Enlightenment “Catholization” of Projective Technology
149
WORKS CITED Apuleius. The Apologia and Florida of Apuleius of Madaura. Trans. H.E. Butler. Westport: Greenwood Press, 1970. Apuleius. The Golden Ass or Metamorphoses. Trans. E.J. Kenney. London: Penguin Books, 1998. Augustine. The City of God against the Pagans. 7 vols. Trans. David S. Wiesen. Cambridge: Harvard University Press, 1968. Barnett, S.J. Idol Temples and Crafty Priests. The Origins of Enlightenment Anticlericalism. New York: St. Martin’s Press, 1999. Beckmann, Johann. A History of Inventions and Discoveries. 4. vols. Trans. William Johnston. London: Longman, Hurst, Rees, Orme, and Brown, 1817. Betz, Hans Dieter. “Magic and Mystery in the Greek Magical Papyri.” Magika Hiera. Ancient Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 244-98. Braarvig, Jens. “Magic. Reconsidering the Grand Dichotomy.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 21-54. Bradley, D.C., G.C. Chang, and R.A. Andersen. “Encoding of Three-Dimensional Structure-from-Motion by Primate Area MT Neurons.” Nature 392 (April 16, 1998): 714-17. Carabelli, Giancarlo. In the Image of Priapus. London: Duckworth, 1996. Cavanaugh, Maureen B. Eleusis and Athens. Documents in Finance, Religion, and Politics in the Fifth Century B.C. Atlanta: Scholars Press, 1996. Christie, James. A Disquistion upon Etruscan Vases. Displaying their Probable Connection with the Shows at Eleusis, and the Chinese Feast of Lanterns, with Explanations of a few of the Principal Allegories Depicted upon them. London: T. Becket. 1806. Christie, James. An Essay on That Earliest Species of Idolatry. The Worship of the Elements. Norwich: Stanhopian Press, 1814. Curl, James Stevens. The Art and Architecture of Freemasonry. An Introductory Study. London: B.T. Batsford, 1991. Damisch, Hubert. L’origine de la perspective. Paris: Flammarion, 1987. David, F.A. Antiquités etrusques, grecques et romaines. Gravees par . . . avec les explications par [Baron] d’Hancarville. Paris: Chez l’Auteur, 1785-1787. Dickie, Matthew W. “The Learned Magician and the Collection and Transmission of Magical Lore.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 163-93. Diderot, Denis. “Eclectisme.” Encyclopédie, ou Dictionnaire Raisonné des Sciences, des Arts et des Métiers. 35 vols. Ed. idem and Jean LeRond d’Alembert. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 1751-1780. Vol 5, 270-93. Diderot, Denis and Jean LeRond d’Alembert. “Eclipse.” Encyclopédie, ou Dictionnaire Raisonné des Sciences, des Arts et des Métiers. 35 vols. Ed. idem. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 1751-1780. Vol 5, 293-98. Du Halde, J.B. Description géographique, historique, chronique, politique, et physique de l’empire de la Chine et de la Tartarie chinoise. Paris: P.G. Lemercier, Imprimeur-Libraire, 1735. Eitrem, Samson. “Dreams and Divination in Magical Ritual.” Magika Hiera. Ancient
150
Barbara Maria Stafford
Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 175-88. Faraone, Christopher A. “The Agonistic Context of Early Greek Binding Spells.” Magika Hiera. Ancient Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 3-32. Fick-Michel, Nicole. Art et mystique dans les Métamorphoses d’Apulée. Paris: Diffusion Les Belles-Lettres, 1991. Fletcher, Richard. The Conversion. From Paganism to Christianity. New York: Henry Holt and Company, 1997. Goethe, Johann Wolfgang von. Italian Journey (1786-1788). Trans. W.H. Auden and Elizabeth Mayer. San Francisco: North Point Press, 1982. Goldstein, E.B. Sensation and Perception. Pacific Grove: Brooks/Cole, 1999. Graf, Fritz. La magie dans l’antiquité greco-romaine. Idéologie et pratique. Paris: Les Belles Lettres, 1994. Graf, Fritz. “Magic and Divinations.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 283-98. Hamill, John. The Craft. A History of English Freemasonry. Wellingsborough, Northamptonshire: Crucible, 1986. Holden, Constance. “The Origin of Speech.” Science 303 (February 27, 2004): 1316-19. Irlam, Shaun. Elations. The Poetics of Enthusiasm in Eighteenth-Century Britain. Stanford: Stanford University Press, 1999. Ittershagen, Ulrike. Lady Hamiltons Attitüden. Mainz: Philipp von Zabern, 1985. Jenkins, Ian. “‘Contemporary Minds.’ Sir William Hamilton’s Affair with Antiquity.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. Jenkins, Ian. “Talking Stones. Hamilton’s Collection of Carved Gems.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. Johnson, Samuel. A Dictionary of the English Language. London: Knapton, 1755. Johnston, Sarah Iles. “Songs for the Ghosts.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. Kahng, Eik. “L’Affaire Greuze and the Sublime of History Painting.” Art Bulletin 86 (March 2004): 96-114. Kerényi, Carl. Eleusis and Athens. Archetypal Image of Mother and Daughter (= Bollingen Series, LXV). Trans. Ralph Manheim. New York: Pantheon, 1967. Kircher, Athanasius. Ars Magna Lucis et Umbrae . . . Amsterdam: Ex Officina Janssonio-Wassbergiana, 1671. Krysmanski, Bernd W. Hogarth’s “Enthusiasm Delineated,” Nachahmung als Kritik am Kennertum. Eine Werkanalyse. Zugleich ein Einblick in das satirisch-aufgeklärte Denkens eines “Künstlerrebellen” im englischen 18. Jahrhundert. Hildesheim: Olms, 1996. Lincoln, Bruce. Emerging from the Chrysalis. Studies in Rituals of Women’s Initiations. Cambridge: Harvard University Press, 1981. Lincoln, Bruce. Priests, Warriors, and Cattle. A Study in the Ecology of Religion. Berkeley, Los Angeles, and London: University of California Press, 1981. Lincoln, Bruce. Theorizing Myth. Narrative, Ideology, and Scholarship. Chicago and London: University of Chicago Press, 1999. Lyons, Claire. “The Neapolitan Context of Hamilton’s Antiquities Collection.” Journal of the History of Collections 9 (No. 2, 1997): 229-40.
The Enlightenment “Catholization” of Projective Technology
151
Martin, Luther H. Hellenistic Religions. An Introduction. New York: Oxford University Press, 1987. Murray, Janet H. Hamlet on the Holodeck. The Future of Narrative in Cyberspace. Cambridge and London: MIT Press, 1997. Ndalianis, Angela. Neo-Baroque Aesthetics and Contemporary Entertainment. Cambridge and London: MIT Press, 2004. Ouvaroff, Serge. [Sergei Semenovich, Count Uvarov (1786-1855)]. Essay on the Mysteries of Eleusis. Trans. J.D. Price with observations by James Christie. London: Rodwell and Martin, 1817. Plato. Republic. Indianapolis: Hackett, 1992. Plato. Phaedrus. Indianapolis: Hackett, 1995. Reeves, Eileen. “Old Wives’ Tales and the New World System. Gilbert, Galileo, and Kepler.” Configurations 7 (Fall 1999): 301-54. Simic, Charles. “Difference in Similarity.” New York Review (11 March 2004): 21-23. Small, Jocelyn Penny. “Time in Space. Narrative in Classical Art.” Art Bulletin 81 (December 1999): 563. Sontag, Susan. The Volcano Lover. A Romance. New York: Farrar, Straus, Giroux, 1992. Spivey, Nigel. Enduring Creation. Art, Pain, and Fortitude. Berkeley, Los Angeles, and London: University of California Press, 2001. Stafford, Barbara Maria. Artful Science. Enlightenment Entertainment and the Eclipse of Visual Information. Cambridge and London: MIT Press, 1996. Stafford, Barbara Maria. Visual Analogy. Consciousness as the Art of Connecting. Cambridge and London: MIT Press, 1999. Stafford, Barbara Maria. “Levelling the New Old Transcendence. Cognitive Coherence in the Age of Beyondness.” New Literary History 35.2 (2004): 321-28. Stafford, Barbara Maria.“‘Fantastic’ Images. From Unenlightening to Enlightening Images Meant To Be Seen in the Dark.” Aesthetic Illusion. Theoretical and Historical Approaches. Ed. Frederick Burwick and Walter Pape. Berlin and New York: de Gruyter, 1990. 158-79. Thackray, John. “The Modern Pliny. Hamilton and Vesuvius.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. 65-74. Tumbleson, Raymond D. Catholicism in the English Protestant Imagination. Nationalism, Religion, and Literature, 1660-1745. Cambridge: Cambridge University Press, 1998. Vernant, Jean-Pierre. Myth and Thought among the Greeks. London, Boston, Melbourne, and Henley: Routledge & Kegan Paul, 1983. Vickers, Michael. “Hamilton, Geology, Stone Vases, and Taste.” Journal of the History of Collections 9.2 (1997): 270. Walsh, P.G. The Roman Novel. The “Satyricon” of Petronius and the “Metamorphoses” of Apuleius. Cambridge: Cambridge University Press, 1970. Warren, Rosanna. Departure. New York: Norton, 2003. Wasson, R. Gordon. “The Divine Mushroom of Immortality.”The Flesh of the Gods. The Ritual Use of Hallucinogens. Ed. P.T. Furst. New York: Praeger, 1972. 185200. Wasson, R. Gordon, Albert Hofmann, and Carl A.P. Ruck. Road to Eleusis. Unveiling the Secret of the Mysteries. New York and London: Harcourt Brace Jovanovich, 1978. Witt, R.E. Isis in the Graeco-Roman World. London: Thames and Hudson, 1971.
JAN LAZARDZIG
The Machine as Spectacle: Function and Admiration in Seventeenth-Century Perspectives on Machines In September 1675, during his stay in Paris, young Gottfried Wilhelm Leibniz witnesses an extraordinary spectacle on the bank of the Seine. He sees the presentation of a “machine qui sert à marcher sur l’eau,” a machine for walking on water. Leibniz is deeply impressed and records his inspiration with the imaginative freedom of a dreamer, composing his “Drôle de Pensée, touchant une nouvelle sorte de REPRESENTATIONS.”1 The topic of this “curious idea concerning a new kind of representation” is an “Académie des représentations” or, as Leibniz adds to the title, an “Académie des Ieux.” All kinds of machines, instruments, and mechanisms were to be shown in this academy of games. These impressive performances would stimulate and direct the audience’s imagination and inspire new inventions. The exceptional position of the Artes mechanicae in the sixteenth and seventeenth centuries gave new meaning to the Latin term machina. The rather static order of the machina mundi was replaced by a dynamic theatrum machinarum.2 At the same time, mechanics, i.e., mechanical regulations and functions, became increasingly important for the state and the individual as well as for physics in general.3 Using 1
2
3
Gottfried Wilhelm Leibniz. “Drôle de Pensée, touchant une nouvelle sorte de REPRESENTATIONS.” Sämtliche Schriften und Briefe. Ed. Akademie der Wissenschaften zu Berlin. Berlin: Akademie Verlag, 1970. Berlin: Akademie Verlag, 1970. Series IV, vol. I, 562-68 and notes 694-96. On Drôle de Pensée and Leibniz’s idea of the theater of nature and art, cf. Horst Bredekamp. Die Fenster der Monade. Gottfried Wilhelm Leibniz’ Theater der Natur und Kunst. Berlin: Akademie Verlag, 2004, especially page 45ff. For the medieval and early modern history of the term “machina,” cf. Marcus Popplow. “Die Verwendung von lat. ‘machina’ im Mittelalter und in der Frühen Neuzeit. Vom Baugerüst zu Zoncas mechanischem Bratenwender.” Technikgeschichte 30.1 (1993): 7-26. On the metaphor of the machine as an organic and mechanical background metaphor, cf. Hans Blumenberg. Paradigmen zu einer Metaphorologie. Frankfurt a.M.: Suhrkamp, 1998. 91ff. For the position of Artes mechanicae in the sixteenth and seventeenth centuries,
The Machine as Spectacle
153
machines and nature analogically, isolated mechanical workings, when forced out of nature, could be impressively presented. Windmills and watermills, for example, could display their superhuman powers for obvious use and pleasure. Admirable mechanical operations came to harmonize with standards of wonder and astonishment at the order of things.4 In this article I will argue that the spectacle, i.e., a performance aimed at an audience, is essential to the idea of the machine in the seventeenth century. Relating to the audience is necessary in bridging the gap between illusion and utility, and allows the machine to become an object of admiration and therefore be guaranteed to “function.” Accordingly, the machine denotes a technique of persuasion, defining perspectives and actions. The machines and automatic devices found in books on mining, mills, and machines by authors ranging from Georg Agricola and Salomon de Caus to Georg Andreas Böckler are the starting point of this article. These machines do not allow an exact differentiation between the pragmatic and the imaginative; everything that can be described mechanically seems technically feasible. The descriptions of machines, especially in the Theatrum machinarum literature, thus appear as an artful staging of personal ambition, and as a spectacular outbidding of human capabilities. The reader, witness to this outbidding as well as its critic, is supposed to participate in the miraculous, i.e., automatic working of the machines.5 Secondly, I will show how the universal machinery of the stage of the theater of machines becomes the exemplary pinnacle of baroque machinery. Novelty and curiosity were the most significant attributes for the perception of machines. As realized in theater of machines pro-
4 5
cf. Jutta Bacher. “Artes Mechanicae.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 35-49. This book includes an extensive bibliography. The Max-Planck-Insitute for the history of science in Berlin is currently developing the Archmimedes Project, a digital research library on the developments in the history of mechanics. Numerous sources are now available on the internet. Cf. the online bibliography under http://archimedes.mpiwgberlin.mpg.de. Cf. Lorraine Daston and Katharine Park. Wonders and the Order of Nature, 11501750. New York: Zone Books, 1998. 215ff. Martin Burckhardt writes of the phantasmatic and the real machine and therefore extends his understanding of the machine to symbolic and semantic dimensions. Cf. Martin Burckhardt. Vom Geist der Machine. Eine Geschichte kultureller Umbrüche. Frankfurt a.M.: Campus, 1999. 7-29.
154
Jan Lazardzig
ductions, they put the audience into a clearly conventionalized position of silent admiration.6 Thirdly, I would like to examine some epistemological implications inherent in understanding the machine as spectacle. Within the scope of Cartesian physics and its mechanical principles, machines were able to expose the doings of nature as simple functional connections. On the other hand, nature could appear as a sort of obscured “machine” to be uncovered. Finally, on the basis of Leibniz’s draft of an “Academy of Representations,” I will show how the spectacle of the machine was understood in an alternative way to that of René Descartes. I. The so-called “machine books,” which appeared in quick succession at the turn of the sixteenth to the seventeenth century, offered a collection of depictions and descriptions of practical and imaginary machines and automatic devices.7 Usually these machines were invented, assembled, and explained by architect-engineers.8 The understanding of what distinguishes a machine was largely drawn from the pseudo-Aristotelian 6
7
8
Blumenberg. Paradigmen. 92f. suggests affect-regulating properties especially of the theater machine. More generally, this applies to mechancical pragmatism, as seen in the following example of machine literature. On the genre of “machine books,” cf. Ansgar Stöcklein. Leitbilder der Technik. Biblische Tradition und technischer Fortschritt. Munich: Moos, 1969, especially 27ff. as well as Marcus Popplow. Neu, nützlich und erfindungsreich. Die Idealisierung von Technik in der Frühen Neuzeit. Münster: Waxmann, 1998. He examines the societal and institutional background of technical literature from a broad basis of sources. For a history of the term, cf. Wilhelm Schmidt-Biggemann. “Maschine.” Historisches Wörterbuch der Philosophie. Ed. Joachim Ritter. Darmstadt: Wissenschaftliche Buchgesellschaft, 1980. Vol. 5, 790-802; Karlheinz Jakob. Maschine, Mentales Modell, Metapher. Studien zur Semantik und Geschichte der Techniksprache. Tübingen: Niemeyer, 1991 (on the development of a German technical language), and Popplow. “Verwendung.” Though the machine books have been examined according to the history of the terms, a comprehensive treatment of the iconography is still lacking. Cf. the remark in Jutta Bacher. “Das Theatrum machinarum. Eine Schaubühne zwischen Nutzen und Vergnügen.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 509-18, as well as fundamental considerations in Marcus Popplow. “Maschinenzeichnungen der Ingenieure der Renaissance.” Frühneuzeit-Info 1 (2004): 13-32. For this broad profile of a Vitruvian ideal, cf. Ulrich Schütte. “Architekt und Ingenieur.” Architekt und Ingenieur. Baumeister in Krieg und Frieden. [Exhibit. cat.] Wolfenbüttel: Herzog August Bibliothek, 1984. 18-31.
The Machine as Spectacle
155
Quaestiones mechanicae, rediscovered in the sixteenth century.9 While these writings remained almost unknown in the Middle Ages, they were now translated into several modern languages and became the influential nucleus of technical literature. The second “motor” of this development was Vitruvius’s De architectura (with the tenth book devoted primarily to war machinery), which was translated and annotated around the same time. The pseudo-Aristotelian text defines the machine as something which man wrests from nature and directs against nature to his own advantage.10 He describes the circle as the “first phenomenon” and thus as the basis of nature and the machine. According to the text, the circle unites inertia and motion as well as forward and backward movement. Due to this inclusion of opposites, it is extraordinary and remarkable and the first of all miracles. The machine hides the circle’s capability and shows only its effect, thus appearing as a miracle: Engineers make use of the circle’s basic natural design in building a machine [órganon]. They hide its (propulsive) principle, so that only the miraculous [thaumastón] aspect of this mechanism [mechánema] is visible while its cause remains in the dark.11
Vitruvius’s well-known definition of a machine follows Aristotle and offers a subdivision into lifting machines, pulling machines, and pneumatic machines: “A machine is a continuous material system having special fitness for the moving of weights. It is moved by appropriate revolutions of circles, which by the Greeks is called cyclice cinesis.”12 9
10 11
12
Fritz Krafft offers commentary and an introduction: Dynamische und statische Betrachtungsweise in der antiken Mechanik. Wiesbaden: Steiner, 1970, especially 1 and 137, which argues for the authenticity of the text. On ‘Rediscovery’ and the meaning of the pseudo-Aristotelian mechanica in the fifteenth century, cf. Paul Lawrence Rose and Stillman Drake. “The Pseudo-Aristotelian ‘Questions of Mechanics’ in Renaissance culture.” Studies in the Renaissance 18 (1971): 65-104; Heribert M. Nobis. “Die wissenschaftstheoretische Bedeutung der Quaestiones Mechanicae.” Der Wissenschaftsbegriff. Historische und systematische Untersuchungen. Ed. Alwin Diemer. Meisenheim: Hain, 1970. 47-63 for the importance of the Quaestiones for the machine literature in the sixteenth and seventeenth centuries. Cf. Krafft. Dynamische und statische Betrachtungsweise. 21f. Krafft. Dynamische und statische Betrachtungsweise. 23f. For a discussion of the influence of pseudo-Aristotelian and Vitruvian mechanics on the machine books, cf. Popplow. “Maschinenzeichnungen.” 98ff. After a thorough study of the sources, Popplow comes to the conclusion that Vitruvius’s complex idea of the machine did not yet include the idea of automatic action, present in modern definitions and in the sixteenth and seventeenth century machine books. Ibid. 103. Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard University Press, 1962. Vol. 2, 275.
156
Jan Lazardzig
This definition can already be found in Agricola.13 His mining treatise De re metallica (1556) mentions a very prominent technique in the sixteenth and seventeenth centuries. In his sixth book, he adopts Vitruvius’s classification of machinae. Technically and functionally, however, he assigns completely different devices to these classifications. Here, machina means the mechanization of processes that are driven by waterpower or physical strength.14 Whenever these technical aids or devices “by far exceed all human power,” the standard for this “Zeug” or “Gezeug” (in the 1557 German translation) becomes the surpassing of human capabilities.15 The attribute of artificiality, which Agricola introduces to differentiate between machines and sheer tools, describes the very special artistic craftsmanship used to create a machine. “There is much stuff that lifts mountains and oceans. It comes in many forms, several of them very artificial.”16 It is also connected with the pseudo-Aristotelian differentiation between instrument and machine. Furthermore, it already implies the technically skillful outwitting of nature that Heinrich Zeising, following the Quaestiones mechanicae, expresses in his 1607 Theatrum machinarum: “Because art is able to defeat man/ in those things/ in which man is defeated by nature.”17 In the early Italian machine books by Agostino Ramelli (1588; dated 1620), Fausto Veranzio (1600), Vittorio Zonca (1607), and Giovanni Branca (1629), the amazing activities of the machines are already being explained by the engineers’ ingenium. These books emphasize the engineers’ painstaking and arduous work again and again. It becomes the direct counterpoint to the desired automation of the machines. Veranzio asks his readers: “Why do I have so much trouble and work that is consuming to describe?”18 He answers his own question with an unprecedented list of machines that save work while promising comfort and ease.19 13 14 15 16 17 18 19
Concerning the following explanations of historical terminology, cf. Jakob. Maschine, Mentales Modell. 119, and Popplow. “Verwendung.” 16. Cf. Popplow. “Verwendung.” 17. Georg Agricola. Vom Bergkwerck 12 Bücher . . . Basel, 1557. 72, 89, 146. Ibid. 72. Heinrich Zeising. Theatri machinarum [6 parts]. Leipzig, 1607-1614. Vol. I, part I, 17. Fausto Veranzio. Machinae Novae Fausti Verantii siceni cum declaratione Latina, Italica, Hispanica, Gallica, et Germanica. Venice, [1600]. The seemingly independent work of the machines also changes the definition of work. Work is that which should not be. The ingenium, acquired by the engineers with great difficulty, testifies to this. Only the machines are truly “ingenious,’ or practice effortless activity. Thus, machine literature of the sixteenth and seventeenth cen-
The Machine as Spectacle
157
Fig. 1: Water pump from Heinrich Zeising. Theatri machinarum (Leipzig, 1607-1614).
The machine books attempt to involve the viewer and reader in various ways. Their goal is to substantiate convincingly the ease and effortlessness of mechanical functioning. The viewer in the picture.20 Even the picture program for Zeising’s Theatrum cannot manage without the viewer (fig. 1). However, not only the expert engineer, together with the financier (identifiable as “aristocratic” by his Spanish court dress) inspect the machine. A noble audience appears to enjoy the divertissement and seems to have accidentally noticed the machine. They approach out of curiosity and admire the machine’s performance. The pointing gesture of (mostly male) hands
20
turies could provide an important point of reference for the development of the term ‘genius.’ Cf. Claude Gandelman. “Der Gestus des Zeigens.” Der Betrachter ist im Bild. Kunstwissenschaft und Rezeptionsästhetik. Ed. Wolfgang Kemp. Berlin: Reimer, 1992. 71-93. This text develops an epoch-specific listing of gestures of pointing in fifteenth century Italian painting.
158
Jan Lazardzig
stands out. The hands point not to the machine as a whole but into the machine’s interior, thus emphasizing and explaining the technical apparatus. This pointing gesture links the viewers in the picture to Zeising’s explanations. He uses complicated words to recapitulate the obvious. At the same time the technical explanations are put into the viewers’ mouths, and they model an admiring attitude towards the machine. In the introduction, Zeising already collapses the role of the viewer in the picture with that of the viewer in front of the picture: I hope to please every refined person who looks carfully at these machines and reads their principles and properties. The gentle reader is truly considered . . . as the most blissful person since he can benefit from these inventions for his own purpose with ease and free from worries.21
The accompanying text relates to the illustrations only as an additional verification of a visually evident functional context. Thus, the text makes the dynamic illustration appear even more convincing. The creation of visibility. The frontispiece of Andreas Böckler’s Theatrum machinarum novum invites the viewer to a seemingly effortless divertissement. On the proscenium, Mechanicus and Archimedes seem to lift the stage curtain as well as the mill house facades, organized in perspective behind the curtain. Through the arch of the proscenium, the viewer sees the insides of the machine houses arranged in perspective. The literal cutting away of the machine casing is a technique used to direct the viewer’s eye and can actually be found in almost all machine books. In Ramelli’s Le diverse et artificiose Machine, the walls of the mills open up, as does the ground, allowing a limited view of underground pipe constructions. As is evident from the sawing machine, the views almost randomly offered by a partial glimpse through walls do not “show everything.” In these views, labeling letters reference the explanatory text, and the text returns the reader to the illustrations. To understand, one must gradually penetrate the walls protecting the engineer’s ingenium (fig. 2). In Ramelli, the principle of forceful opening continues with screwing and lifting mechanisms for breaking open enormous city gates and with powerful catapults for breaking through city fortifications. The halving and breaking open of these apparatuses are a product of the machinery that they reveal. The machines presented become a circle of production and destruction. The visibility granted here results from an exposure that embodies the principle of the machine as a tool set against nature. Through the broken walls, the viewer’s gaze itself becomes a mechanical, destructive gaze. 21
Zeising. Theatri machinarum. Vol.1, part 1. “Vorrede.”
The Machine as Spectacle
159
Fig. 2: Sawing machine from Agostino Ramelli. Schatzkammer mechanischer Künste (Leipzig, 1620).
The dynamic processes between picture and viewer sketched here contribute to overcoming the inertia of mechanical devices and to realizing the idea of personal ambition. Thus, they strongly emphasize the spectacular character of these machines. The machines are not only unique technical examples, but are also surrounded by an aura of drama and staging obviously directed at an audience. In this sense, they are actually theater machines. II. The word “machine” in European languages derives from Latin. In seventeenth century lexicography, the term describes both machines and theater machines. Therefore the theater machine directly absorbed mill-
160
Jan Lazardzig
ing and lifting machines, pneumatic and automatic machines. Antoine Furetière, in his 1690 Dictionnaire, gives an overview of the term machine.22 Proceeding first from the Vitruvian definition, he takes his examples almost exclusively from theater machinery, which becomes his model for explaining all relevant features of the machine. He borrows his definition word-by-word from the work of Claude-François Menestrier, a Jesuit theoretician and emblematic. “In general one calls everything a machine that draws its movement only from human craftsmanship, such as theater sets, namely the wagons, the clouds, the ships.”23 After Agricola, superhuman exertion had become the sign of the machine. The machine books stage this impressively in the wheel gearings in the mills. For Furetière, the flying devices of the theatrical universal machinery perfectly realize this exertion. The flying machines allow man “to do things that exceed his own powers, such as flights, landings, etc.”24 They most clearly showed the superhuman qualities (symbolically also a divine attribute) inherent in machines. These machines triggered astonishment and surprise in the audience, according to Furetière precisely because they did not reveal the secret of how they functioned. He includes an “invention for changing decoration, for flying through the air, for the movement of animals and other technical skills that surprise and entertain the viewers, as they do not know their secrets.”25 Astonishment and amazement at the extraordinary power of the machine are linked to the artful concealment of the technical apparatus. In this they resembled automatic machines that seemed to move independently, for example machines powered by wind and water (or by animals) such as pumps, mills, hydraulic and pneumatic machines, which functioned without constant human intervention.26 22 23 24
25 26
Antoine Furetière. Dictionnaire universel . . . 3 vols. Den Haag and Rotterdam, 1690. Vol. 2, entry for “machine.” Claude-François Menestrier. Traité des Tournois, Ioustes, Caroussels, et autres Spectacles publics. Lyon, 1669. 77. Furetière. Dictionnaire universel, “machine.” These theater machines are the inspiration for a series of others, also meant to replace (and multiply) human physical power. On the theater machines of the seventeenth and eighteenth centuries, cf. Madeleine Horn-Monval. “La grande machinerie théâtrale et ses origines.” Revue d’histoire du théâtre 4 (1958): 291ff.; Margret Dietrich. “Vom Einfluss der Mathematik and Mechanik auf das Barocktheater.” Sonderabdruck aus dem Anzeiger der phil.-hist. Klasse der Österreichischen Akademie der Wissenschaften 107.1 (1970): 7-22. Furetière. Dictionnaire universel, “machine.” Moreover, a machine could mean something serious and important, a weighty matter. “MACHINES, se dit aussi des choses pesantes & difficiles à remuer, à transporter.” Furetière. Dictionnaire universel, “machine.”
The Machine as Spectacle
161
He describes watches as “the most beautiful machines ever to be invented.”27 Even as late as Denis Diderot’s Encyclopédie, the machine primarily functioned as a spectacle.28 As such, theater machines count as evidence of a baroque (theater) aesthetic.29 “Everything which is done by machines always seemed to be admirable, splendid and surprising.”30 Menestrier thus describes the effect of the machines on the viewers. In Les Caractères, Jean de la Bruyère gushes that the machines enhance the fiction and support the viewers’ sweet illusion, which creates the entire pleasure of theater.31 And, François Hédelin Abbé d’Aubignac, who clearly expressed reservations concerning the Piéces en machine because of the common failure to work the machines reasonably, allows their use in theater.32 He gives precise directions for the aesthetic employment of machines in theater to produce “admiration” and “étonnement” (astonishment).33 According to the Abbé Michel de Pure in his Idée des spectacles anciens et nouveaux (from 1668), two distinct forms of beauty affect the senses. The outward beauty of the machines captivates the view (“elle dérobe aux yeux”) by their surprising novelty, as they oppose what the viewer considers (obviously) possible or reasonable. Secondly, the invention and performance of new movements and activities reveals the soul, so to speak, of this ‘natural magic,’ (“enchantement naturel”). This hidden 27
28
29
30 31 32 33
Furetière. Dictionnaire universel, “machine.” Peter Rück. “Die Dynamik mittelalterlicher Zeitmaße und die mechanische Uhr.” Die Mechanik in den Künsten. Studien zur ästhetischen Bedeutung von Naturwissenschaft und Technologie. Ed. Hanno Möbius and Jörg Jochen Berns. Marburg: Jonas Verlag, 1990. 17-30. This underlines the paradigmatic importance of the clock for mechanics between 5001500. The analogy of a mechanical clock with wheels and the movement of celestial bodies was first proposed by Nicole Oresme around 1350. The clock with wheels first became a characteristic of “orderly functioning.” Cf. Popplow. “Verwendung.” 15, 20. Thus, in Diderot’s Encyclopédie of 49 large format panels on the theme of Machines de théâtre, 88 fall into the area of Spectacle. While the entire field of agriculture, including the images of water and windmills, oil and wine presses only comes to 85 panels. Cf. Andreas Gipper. Wunderbare Wissenschaft. Literarische Strategien naturwissenschaftlicher Vulgarisierung in Frankreich. Munich: Fink, 2002. 344. Cf. for example Thomas Edward Lawrenson. The French Stage and the Playhouse in the XVIIth Century. A Study in the Advent of the Italian Order. New York: Ams Press, 1986. 150ff. Menestrier. Traité des Tournois. 141 Jean de la Bruyère. Œuvres complètes. Ed. Julien Benda. Paris: Gallimard, 1951. 79. François Hédelin d’Aubignac. La pratique du théâtre . . . Amsterdam, 1715. 320f. and 322f. Ibid. 324ff.
162
Jan Lazardzig
capability moves elements appropriately according to the occasion and the will of the machinist. The poet should primarily make use of the initial beauty of the spectacle. It is his task to so enthrall the senses of the viewers that they cannot even formulate a question regarding the technical means of the magic. It must please the senses and those eyes, whose gaze does not reach beyond visible things and who do not think it worth the effort to ponder the functioning of things or the reason for their pleasure which cajoles them.34
In the above, the calculated dramatic employment of theater machines resembles the young actress whose task was described by de Pure as disarming the intellect: “Convincing the mind is simple, once the senses are satisfied.”35 The change from relatively static uniform decorations by Baldassare Peruzzi, Sebastiano Serlio, and Andrea Palladio towards moving theater machinery (Periaktoi by Antonio da Sangallo, machines and divine appearances by Bernado Buontalenti) accompanied a sequence of mechanical innovations in the eighties and nineties in sixteenth century Italy, above all from the engineering city of Ferrara. Therefore, it is no coincidence that the builder of the Teatro Farnese in Parma, Giovanni Battista Aleotti, was one of the most significant hydraulicists of his time. (His Idrologia overo Vaso delle Scienze et Arte delle Acque del ben regulare le acque is preserved in five manuscript volumes.) The three generations of machinists from Italy that revolutionized Parisian theater in the seventeenth Century, Tomaso Francini, Giacomo Torelli and Gaspare Vigarani, were originally engineers, architects, mathematicians, mechanics, and hydraulicists. Often they combined all these skills, as in the case of Torelli, one of the most celebrated machinists of the seventeenth century.36 From changing scenery to artful apparatuses for creating illusions to controlled devastation by the four elements37 to spectacular machine-supported emanations of the godly, the dense sequence of technical innovation accelerated and influenced the way in which theater machinery fulfilled and formed the baroque imagination. This process is significantly tied to the conditions surrounding the performances of the theater of machines. The fame of the theater of ma34 35 36 37
Michel de Pure. Idée des spectacles anciens et nouveaux. Paris, [1668]. 301ff. Ibid. 170. Cf. Per Bjurström. Giacomo Torelli and Baroque Stage Design. Stockholm: Almqvist & Wiksell, 1961. Cf. the relevant chapters in Nicola Sabbattini. Pratica di fabricar scene, e machine ne’ teatri . . . Ravenna, 1638, Secondo libro, Cap. 51. “Il Vento come si finga;” Cap. 52. “Come si poßano fingere I Lampi;” Cap. 53. “I Tuoni come si fingano.”
The Machine as Spectacle
163
chines was mostly based on on-off performances for a select noble audience accompanying a special event, an engagement, a wedding, or a peace treaty.38 The performance lived on in the written descriptions of the festivities. The machines thus described in the Parisian journals filled the heads of those who could not be present. Exceptional machine spectacles, which satisfied the public’s lust for sensation, could become a life-long attribute of the machinist.39 The spectacular and novel qualities of the machines were verified by the written reports and also became the convention for speaking about the machines. The countless Pièces or Tragédies en machine, written especially for spectacle apparatuses (d’Aubignac warns precisely against this development) in the foundation phase of the French opera, constituted a genre of their own, if only for a short time.40 The accounts of the performances that surrounded this genre, above all the monthly Mercure galant, whose editor Donneau de Visé himself wrote a machine piece, euphorically show the use and the effect of the theater machines.41 The theater performances of the famous Marquis de Sourdéac, whose palace was close to Leibniz’s Parisian hôtel, were described as follows. “He created something so overwhelming, of such beauty and novelty that people came from all four corners of the world to admire it.”42 His machines were “built with all imaginable accuracy.”43 The theater equipped by Sourdéac for the Pièce à machine Circè (performed at the time of Leibniz’s sojourn in Paris) was described by the Gazette d’Amsterdam as “surpassing all power of imagination.”44 Countless other examples could be named. They all contribute not only to the discussion of machines, but also to an interpretation of the conventionalization expressed in this understanding of the machines. This conventionalization totally ignores moments of machine dysfunction and clearly shows the material border of the mechanization of theater. For example, the Théâtre des Machines 38 39
40
41 42 43 44
For examples cf. Margarete Baur-Heinhold. Theater des Barock. Festliches Bühnenspiel im 17. und 18. Jahrhundert. München, 1966. As, for example, in the case of the Marquis de Sourdéac, who was obsessed with theater and a gifted machinist. Cf. V. and M. Delavigne. “Un grand Seigneur au XVIIe siècle. Le Marquis de Sourdéac.” La Revue Hebdomadaire 20 (1911): 450-83. Thomas Corneille wrote numerous of these pieces. Cf. Margret Dietrich. “Der Barocke Corneille. Ein Beitrag zum Maschinen-Theater des 17. Jahrhunderts.” Maske und Kothurn 4 (1958): 199-219, 316-45. Jean Moynet. L’envers du théatre. Machines et décorations. Béziers: Hachette, 1990 [facsimile of the edition Paris, 1873]. 18. Mercure galant. Paris, 1673. Vol. III, 341f. Ibid. 337 Gazette d’Amsterdam (February 14, 1675).
164
Jan Lazardzig
(Théâtre des Tuileries), which Gaspare Vigarani equipped in 1662 took on such gigantic proportion that the noise produced by the machines prohibited music and speech.45 Only occasional hints suggest that the risky techniques also provided opportunity for mockery and laughter if they failed. In his poetics of the early period of the theater of machines in Paris Mesnadière writes, “It is true, that these technical skills sometimes please the eye, but one must also consider their disadvantage of often causing laughter in simple people – even in the older ones – if they do not function according to the criteria of the stage.”46 III. One result of Cartesian physics was a clarification of the method and function of the cosmological metaphor of the machine – as established in Johannes Kepler’s change in metaphors from the cosmos instar divini animalis to the cosmos instar horologii (1615) – not only in its astronomical, but also in its mechanical context.47 He gave the metaphysical conditions through the methodic separation of res extensa and res cogitans, making a mechanical and geometric explanation of complete physics possible. From this point on, the attribute of naturalness could describe machines as well as plants. Nicolas Joseph Poisson, editor of the Traité de la Méchanique, included in the Discours de la méthode, gives the following explanation: One must be careful not to let oneself be deceived regarding the word “mechanics,” which refers not only to the science of machine construction or a knowledge of its components. This term also includes the different ways in which the body moves according to various irrefutably fixed natural laws.48
In his text Traité de l’homme, Descartes offers the first consequential application of mechanics to anthropology. In a technomorphous interpretation of the human body he draws on the pipe systems of the artificial grottos and fountains of royal gardens to illustrate the structure of the nervous system.49 In this he calls upon personal experiences (Fontain45
46 47 48 49
Pure. Idée des spectacles. 313. The stage had immense dimensions: 46m long, 20.6m wide and 17m high. The floor for the flying machine again had a height of 6.66m, the stage for the hell machines a depth of 5.28m. The hall could hold five to six thousand people. Cf. Horn-Monval. “La grande machinerie.” 306. Jules de la Mesnadière. La Poëtique. Paris, 1640. 418. Schmidt-Biggemann. “Maschine.” 792. René Descartes. Traité de la Mechanique . . . Paris, 1668. 18. Cf. René Descartes. “The Treatise on Man.” The Philosophical Writings of Des-
The Machine as Spectacle
165
Fig. 3: Representation of a grotto from the second book of Salomon de Caus. Les raisons des forces mouvantes (Frankfurt, 1615).
bleau’s garden) and also written descriptions (the diary of Michel de Montaigne’s Italian journey in 1580 and 1581).50 Descartes illustrates the animated spirits (esprits animaux), which should guarantee an interaction between body and soul, with the power of water, which keeps a mill or an automatic organ running. At the same time, water functions to transport the animated spirits. Descartes interprets the pipes that embody the mechanical principle as the infrastructure by which the spirits move in the body. The physical process of perception becomes a machine spectacle borrowed from Salomon de Caus’s machine book (fig. 3):
50
cartes. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. I, 99, 101f. The parallels with the machines presented by Salomon de Cause, especially his Grottos, are unmistakable. Cf. Salomon de Caus. Von gewaltsamen Bewegungen . . . Hanover: Vincentz, 1977 [facsimile of the edition Frankfurt, 1615]. Book I, Problemata XXVII. Cf. René Descartes. Œuvres de Descartes. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1996. Vol. 11, 212ff. for further examples. I thank Ludger Schwarte for this reference.
166
Jan Lazardzig
External objects, which by their mere presence stimulate [the body machine’s] sense organs and thereby cause them to move in many different ways depending on how the parts of its brain are disposed, are like visitors who enter the grottos of these fountains and unwittingly cause the movements which take place before their eyes. For they cannot enter without stepping on certain tiles which are so arranged that if, for example, they approach a Diana who is bathing they will cause her to hide in the reeds and if they move forward to pursue her they will cause a Neptune to advance and threaten them with his trident; or if they go in another direction they will cause a sea-monster to emerge and spew water onto their faces; or other such things according to the whim of the engineers who made the fountains. And finally, when a rational soul is present in this machine it will have its principal seat in the brain, and reside there like the fountain-keeper who must be stationed at the tanks to which the fountain’s pipes return if he wants to produce, or prevent, or change their movements in some way.51
Because Descartes employed the technically elaborate machine spectacle of the grotto as an allegory of the human apparatus of perception, the unique position of the affect of admiration, which Descartes explains in Passions de l’âme (1650), seems to be prefigured in a qualitative way.52 Descartes interprets admiration as the pre-reflective initial disposition of every act of perception. It regulates what our attention is initially drawn to, and therefore what is perceived. By means of “a movement of the spirits, which is brought about by the impression,” admiration situates each impression.53 Moderation and balance are needed, in order not to let silencing slip into astonishment (étonnement), in order to allow the very reflection that follows the pause. Therefore, the economics of attention that Descartes places in the affect of admiration seem at first different from the conventions for the reception of spectacle machinery. While spectacle machinery aims to prevent or at least to guide reflection by means of overwhelming sensory impressions, so admiration should allow reflection and assessment of the observed. Yet both func51
52
53
Descartes. “The Treatise on Man.” 101f. Cf. also Fernand Hallyn. “Aspects de la problématique de l’illusion chez Descartes.” L’illusion au XVIIe siècle (= Littératures classiques, vol. 44). Ed. Patrick Dandrey and Georges Forestier. Paris: Champion, 2002. 284-304. This article suggests that Descartes always regards machines as producing a spectacle. In Réflexions critiques sur la poésie et la peinture (Paris, 1719), Abbé DuBos suggests the limits of using mechanics as a model for the functioning of perception apparatuses. Cf. René Descartes. “The Passions of the Soul.” The Philosophical Writings of Descartes. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. I, 353f. For Descartes it is a question of the preservation of power. The power of the impression which causes astonishment should be retained, and that is why there is the effect of silencing.
The Machine as Spectacle
167
tion within the same parameters. Their yardstick is the new and the surprising.54 An example of this form of philosophical astonishment is even found in the Meditationes. Descartes describes a moment of confusion that seizes him as people walk by his window. This teichoskopie placed before the inner eye of the reader stages his sudden insight into the susceptibility of the senses as astonishment at the possibility of perfect human machines: But then if I look out of the window and see men crossing the square, as I just happen to have done, I normally say that I see the men themselves . . . Yet do I see any more than hats and coats that could conceal automatons? I judge that they are men. And so something which I thought I was seeing with my eyes is in fact grasped solely by the faculty of judgment which is in my mind.55
The play of costumed machines framed by the window removes the reader to a world of deception, to the theater. The role of the first-person narrator, therefore of the perceiver, corresponds to the master of machinery. This figuration is also found in the summary of Cartesian epistemological philosophy in which Bernard Le Bovier de Fontenelle (1657-1757), later president of the French Academy of Sciences, likens this figuration to an opera première. His Entretiens sur la pluralité des mondes (1686) begins an astronomy lesson for a noble woman with the following assessment: “All philosophy,” I told her, “is based on two things only: curiosity and poor eyesight . . . I have always thought that nature is very much like an opera house. From where you are at the opera you don’t see the stages exactly as they are; they’re arranged to give the most pleasing effect from a distance, and the wheels and counter-weights that make everything move are hidden out of sight. You don’t worry, either, about how they work. Only some engineer in the pit, perhaps, may be struck by some extraordinary effect and be determined to figure out for himself how it was done. That engineer is like the philosophers. But what makes it harder for the philosophers is that, in the machinery that Nature shows us, the wires are better hidden – so well, in fact, that they’ve been guessing for a long time at what causes the movements of the universe . . . ”56
54
55
56
Hans Georg Nicklaus. “Opern des Barock als technisches Spektakel.” Frühneuzeit-Info 1 (2004): 42f. Here, the affect of admiration is bound to the fascination with the technical. René Descartes. “Meditations on First Philosophy.” The Philosophical Writings of Descartes. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. II, 21. Bernard Le Bovier de Fontenelle. Conversations on the Plurality of Worlds. Trans. H.A. Hargreaves. Berkeley and Los Angeles: University of California Press, 1990. 11. Cf. Gipper. Wunderbare Wissenschaft. 134ff. on this comparison, which evident-
168
Jan Lazardzig
If the perspectival space of baroque opera ideally expresses the architecture of Cartesian epistemological philosophy,57 then turning away from mechanics and a different concept of the spectacle underlines Leibniz’s critique of Descartes. IV. Leibniz, in the course of his analysis of the Cartesian teaching on substances, rejects the analogy of the theater of machines from Fontenelle’s Entretiens in his annotations to his article Neues System contributed to the Journal des Savants published in Paris in 1695. According to Leibniz, the Cartesians imagine, “that the difference between their machines and ours is only a difference between large and small.” This misconception produced Fontenelle’s response, that on close inspection nature appears less wonderful than we had thought, it being only something like a craftsman’s window display. I think that this gives an inappropriate and unworthy idea of nature, and that it is only my system which shows the true and immense distance there is between the least productions and mechanisms of divine wisdom and the greatest masterpieces produced by the skill of a limited mind – a difference which is not merely one of degree, but one of kind.58
He explains the incompatibility of divine machines and the human machines that produce divinity by the example of two clocks. He writes, “Imagine two clocks, or two watches which always tell exactly the same time.” And in a later contribution to the Journal des Savants (1696): “This can be done in three ways. The first is by the mutual influence of one clock on the other; the second, by the attentions of a man who looks after them; the third, by their own accuracy.”59 The complete correspondence between the two clocks is neither due to an inexplicable
57
58
59
ly refers to Phaéton, a tragédie lyrique by Philippe Quinault, performed three years before. Cf. Angelica Horn. “Das Experiment der Zentralperspektive. Filippo Brunelleschi und René Descartes.” Descartes im Diskurs der Neuzeit. Ed. Wilhelm Friedrich Niebel and Herbert Schnädelbach. Frankfurt a.M.: Suhrkamp, 2000. 9-32. This translates the Cartesian cogito into central perspectival structures. Gottfried Wilhelm Leibniz. “New System of the Nature of Substances and their Communication, and of the Union which Exists between the Soul and the Body.” Philosophical Texts. Ed. and trans. R.S. Woolhouse and Richard Francks. Oxford: Oxford University Press, 1998. 148. Gottfried Wilhelm Leibniz. “Extract from a Letter Written by Monsieur Leibniz about his Philosophical Hypothesis.” Ibid. 192.
The Machine as Spectacle
169
and therefore wonderful “reciprocal influence,” as testified by Christiaan Huygens, nor provided by divine intervention. According to Leibniz, both would be assuming a “deus ex machina.”60 Instead, he traces the correspondence between the clocks to a common cause and rejects the allegorical use of machines and bodies in favor of an order that regulates the independent interaction of the elements with each other, like “a spiritual or formal automaton . . . which is endowed with a share of reason.”61 This machine could independently produce that which strikes the soul. It would contain inner tendencies, for example individual earlier experiences would be present in this machine. And because these tendencies are the starting conditions for later perceptions, future perception is contained in them: “For in conformity with a law of order which exists in perceptions as much as in motions, each preceding perception influences succeeding one.”62 “Perception, and everything that depends on it,” as he writes later in the Monadologie, “is inexplicable by mechanical principles, by shapes and motions, that is.” Imagine there were a machine which by its structure produced thought, feeling, and perception; we can imagine it as being enlarged while maintaining the same relative proportions, to the point where we could go inside it, as we would go into a mill. But if that were so, when we get in we would find nothing but pieces which would push one against another, and never anything to account for perception.63
Here, the monad, itself, described as a “kind of divine machine or natural automaton,” replaces perceptive apparatuses working according to mechanical principles.64 It is not subject to material construction, but instead is metaphysically described as a simple or complex monad. Therefore, according to Leibniz, natural machines and thus all organic bodies are vastly superior to artificial automatons. A machine built by humans, “is not a machine in every one of its parts . . . But nature’s machines – living bodies, that is – are machines even in their smallest parts, right down to infinity.”65 With this diagnosis, Leibniz aims at the logic of representation of those mechanical machines created for the illusion of independent movement. His idea of an extensive machine, an ‘automaton-monis60
61 62 63 64 65
Ibid. 192 and Gottfried Wilhelm Leibniz. “A Letter from M. Leibniz to the Editor, Containing an Explanation of the Difficulties which M. Bayle Found with the New System of the Union of the Soul and Body.” Philosophical Texts. 205. Leibniz. “New System.” 151. Leibniz. “Explanation of Difficulties.” 206. Leibniz. “Monadology.” Philosophical Texts. 270. Ibid. 277. Ibid.
170
Jan Lazardzig
mus,’ springs from the representative space of the Cartesian theater of machines and makes clear the difference between machine and automaton, between mechanics and organization.66 The spectacle of this Espèce d’un Automate was already sketched in Leibniz’s early draft of ‘Academy of Representation.’ Leibniz’s text is important not only as the first culmination of his model for society.67 We also find his earliest reflections on the theory of games,68 comments on perspective study, as well as the collection and exhibition principles he followed his entire life.69 The space outlined in Leibniz’s Drôle de pensée is a subjective space of experience, encounters, and curiosities. The city topography of Leibniz’s Paris70 and numerous spectacular forms of performance such as comedies, horse ballets, fireworks, magic lantern presentations, lotteries and games of chance, anatomical demonstrations, automaton and machine displays, public experiments, cabinets of curiosities, an office of registration for inventors, and much more blend together in this space.71 An “academy of games” and a “shadow theater” are extensively addressed as places for strategic deception and illusions of perspective. The literal multifacetedness of demonstrations, spectacles, and plays, which were strung together without a serial organizing principle, made it difficult to find a central location or building for his academy. “It could exist in multiple buildings in various districts of the city or rather in different rooms such as royal shops in which private individuals rent 66
67
68
69
70
71
He writes retrospectively about his departure from mechanics. “I am as ready as anyone to do justice to the moderns; nevertheless I think they have carried reform too far, among other things in conflating natural things with artificial ones, through not having sufficiently grand ideas of the majesty of nature.” Leibniz. “New System.” 148. Ines Böger’s book informs us extensively about Leibniz’s model for society. Ines Böger. ‘Ein Seculum . . . da man zu Societäten Lust hat.’ Darstellung und Analyse der Leibnizschen Sozietätspläne vor dem Hintergrund der europäischen Akademiebewegung im 17. und 18. Jahrhundert. 2 vols. Munich: Utz, 1997. On his sojourn in Paris, cf. especially vol. 1, 96ff. On the essays of Leibniz’s theory of games, cf. Marc Parmentier, ed. L’estime des apparences. 21 manuscripts de Leibniz sur les probabilités, la théorie des jeux, l’espérance de vie. Paris: Vrin, 1995. On Drôle de Pensée, cf. especially 19ff. Cf. Wilhelm Ennenbach. “Über eine öffentliche Einrichtung zur Vorführung, Lagerung und Erfassung technischer Objekte.” Neue Museumskunde 24.2 (1981): 103-08 as well as Bredekamp’s extensive bibliography in Die Fenster der Monade. Cf. Paul Wiedeburg. Der junge Leibniz. Das Reich und Europa. Part II: Paris. Vol. I: Europäische Politik. Wiesbaden: Steiner, 1970. 610ff. and vol. III (notes), 293ff. At the beginning of the text he suggests high ranking and wealthy machine devotees as financial partners for these insitutions.
The Machine as Spectacle
171
rooms to display curiosities.”72 The mere number of institutions to be included, such as Colbert’s four nation collegium, the Théâtre du Marais, a multitude of gambling houses, sporting venues, art galleries, etc. did not allow a single location. Moreover, the “Academy of Representation” proliferated into a multiplication of perspectives, opinions and insights that themselves led to new inventions. This enterprise would be useful for the private and the public sphere. It would open the public’s eyes, encourage new inventions, give them pleasant insights and would inform the world of useful or intelligent novelties. All those who had a new invention or an intelligent design could gather here, earn a living, and make their inventions famous in order to make a profit.73 The interrelationship between excitement and invention would guarantee the self-sustaining atmosphere of the academy. In addition, Leibniz repeatedly includes considerations aimed at securing financial autonomy. Often no costs would ensue, as only with a certain payment would others be allowed to display in the building of the academy. And those who stayed in the academy would even make a profit. One would have no expenses.74
Several times, Leibniz thematizes the social rules, which would ensure the survival of the facilities. He considers a developed surveillance system for the many Académies de jeu.75 These are the location of parties, gambling, and also masques. He hoped this surveillance would promote the well-being of the state.76 The emphasis on military drills and athletic competitions reveals his plans for physical as well as intellectual education.77 In the otherwise staccato listings of Drôle de Pensée, the testimony to games of chance and shadow theater stand out. In a significant state of 72 73 74 75
76
77
Leibniz. “Drôle de Pensée.” 565. Ibid. 565. Ibid. 566. Académies de jeu were the establishments for games of chance in Paris in the seventeenth and eighteenth centuries. Leibniz must mean both the place of gambling as well as an academy, or a school of games. Cf. Furetière. Dictionnaire universel. Vol. II, entry for “jeu.” On the culture of play, cf. Francis Freundlich. Le monde du jeu à Paris 1715-1800. Paris: Albin Michel, 1995. Athanasius Kircher and Georg Philipp Harsdörfer offer an illustration of immense earpieces as a room surveillance instrument. See the reproductions in Jörg Jochen Berns’s contribution to this volume. Here considerations which Leibniz wrote about in his Altdorfer Dissertation (written in 1666 and published anonymously in 1667) are important. He writes of the “gradual establishment of habit.” Cf. Hubertus Busche. Leibniz’ Weg ins perspektivische Universum. Eine Harmonie im Zeitalter der Berechnung. Hamburg: Meiner, 1997. 179. Busche identifies the Nova methodus discendae docendaeque jurisprudentiae (1667) as a sketch of a “general philosophy of education.” Ibid. 172.
172
Jan Lazardzig
indecision, Leibniz wonders if cheats should be allowed into theaters or not. Finally he comes to the conclusion that gambling must remain the nucleus of the entire project, because in play even deception entails a healing effect. Games would provide the best excuse in the world to begin such a useful thing for the public, since one needs to fool people, to profit from their weakness and to deceive in order to heal. There is nothing better than using machines for introducing wisdom. This truly means miscere utile dulci and to make medicine from poison.78
The dependence on the miscere utile dulci (Ars Poetica, V. 343,79 originally from Horace but later developing into a topos in baroque poetry) emphasizes the stubborn position of deception and illusion in the ‘Academy of Representation.’ Horace’s formulation aims at a balanced relationship between reality and fiction in poetry, “ficta voluptatis causa sint proxima veris” (Fictions that aim to please should come close to reality. Ars Poetica, V. 338). Yet Leibniz promotes illusion (in the sense of the Latin stem illusio: deception and illudere: to play with/for) to the principle of his academy. Sense is articulated only in the accumulation of illusion.80 The ‘Academy of Representation’ feeds the world of events with countless perceptions. At the same time these lead to the development of abilities and resemble organic growth. The independence of the cognitive process striven for in the structure of the academy could not function with enclosure or fencing in. Instead it required addition and extension. Actually, the mechanical is less of a standard for this enterprise than automatizing in the sense of a creative act of “becoming organic.” The theater of the arts, which Leibniz develops in Drôle de Pensée, “an infinite machine of which every part is a machine,”81 departs from an illusion aimed at functional mechanical equality between nature and the machine. This theater strips and exposes nature without coming to 78 79
80
81
Leibniz. “Drôle de Pensée.” 567. On the role of this topos in the sixteenth and seventeenth centuries, cf. Jean-Marie Piemme. “L’utile dulci ou la convergence des nécessités. Recherches historiques sur les causes de l’adoption de la règle scaligérienne de l’utilité, par les théoriciens de 1630.” Revue d’histoire du théâtre 2 (1969): 118-33. This article emphasizes the change in meaning, the strengthening and the ‘becoming the rule’ of the phrase from Horace to Julia Caesar Scaliger. In reference to theater, Bonnefoy characterized baroque ontology as a blending of illusion and disillusion for the production of being. Cf. Yves Bonnefoy. Rome, 1630. L’Horizon du premier baroque. Paris: Flammarion, 1994. 151. Gilles Deleuze. The Fold. Leibniz and the Baroque. Trans. Tom Conley. Minneapolis: University of Minnesota Press, 1992. 124.
The Machine as Spectacle
173
its end, without reaching the wall behind the stage of the Cartesian opera house. The figure of the harlequin, which Leibniz describes in Nouveaux Essais, becomes relevant here. The harlequin is a stage figure who endlessly removes seemingly infinite costume layers. Yet here, being conscious of the necessary end, he marks the limits of the comparison, as “nature’s artifice is of an entirely different order of subtlety.”82
Translation: Laura Bohn, Martin Wittenberg WORKS CITED Agricola, Georg. Vom Bergkwerck 12 Bücher . . . Basel, 1557. Aubignac, François Hédelin d’. La pratique du théâtre . . . Amsterdam, 1715. Bacher, Jutta. “Artes Mechanicae.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 35-49. Bacher, Jutta. “Das Theatrum machinarum. Eine Schaubühne zwischen Nutzen und Vergnügen.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 509-18. Baur-Heinhold, Margarete. Theater des Barock. Festliches Bühnenspiel im 17. und 18. Jahrhundert. München, 1966. Bjurström, Per. Giacomo Torelli and Baroque Stage Design. Stockholm: Almqvist & Wiksell, 1961. Blumenberg, Hans. Paradigmen zu einer Metaphorologie. Frankfurt a.M.: Suhrkamp, 1998. Böger, Ines. ‘Ein Seculum . . . da man zu Societäten Lust hat.’ Darstellung und Analyse der Leibnizschen Sozietätspläne vor dem Hintergrund der europäischen Akademiebewegung im 17. und 18. Jahrhundert. 2 vols. Munich: Utz, 1997. Bonnefoy, Yves. Rome, 1630. L’Horizon du premier baroque. Paris: Flammarion, 1994. Bredekamp, Horst. Die Fenster der Monade. Gottfried Wilhelm Leibniz’ Theater der Natur und Kunst. Berlin: Akademie Verlag, 2004. Burckhardt, Martin. Vom Geist der Machine. Eine Geschichte kultureller Umbrüche. Frankfurt a.M.: Campus, 1999. Busche, Hubertus. Leibniz’ Weg ins perspektivische Universum. Eine Harmonie im Zeitalter der Berechnung. Hamburg: Meiner, 1997. Caus, Salomon de. Von gewaltsamen Bewegungen . . . Hannover: Vincentz, 1977 [facsimile of the edition Frankfurt, 1615]. Daston, Lorraine and Katharine Park. Wonders and the Order of Nature, 1150-1750. New York: Zone Books, 1998. Delavigne, V. and M. “Un grand Seigneur au XVIIe siècle. Le Marquis de Sourdéac.” La Revue Hebdomadaire 20 (1911): 450-83. 82
Gottfried Wilhelm Leibniz. New Essays on Human Understanding. Ed. and trans. Peter Remnant and Jonathan Bennet. Cambridge: Cambridge University Press, 1996. 214.
174
Jan Lazardzig
Deleuze, Gilles. The Fold. Leibniz and the Baroque. Trans. Tom Conley. Minneapolis: University of Minnesota Press, 1992. Descartes, René. Traité de la Mechanique . . . Paris, 1668. Descartes, René. “The Passions of the Soul.” The Philosophical Writings of Descartes. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. I, 325-404. Descartes, René. “The Treatise on Man.” The Philosophical Writings of Descartes. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. I, 99-108. Descartes, René. “Meditations on First Philosophy.” The Philosophical Writings of Descartes. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. II, 1-62. Descartes, René. Œuvres de Descartes. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1996. Dietrich, Margret. “Der Barocke Corneille. Ein Beitrag zum Maschinen-Theater des 17. Jahrhunderts.” Maske und Kothurn 4 (1958): 199-219 and 316-45. Dietrich, Margret. “Vom Einfluss der Mathematik and Mechanik auf das Barocktheater.” Sonderabdruck aus dem Anzeiger der phil.-hist. Klasse der Österreichischen Akademie der Wissenschaften 107.1 (1970): 7-22. DuBos, Jean Baptiste. Réflexions critiques sur la poésie et la peinture. Paris, 1719. Ennenbach, Wilhelm. “Über eine öffentliche Einrichtung zur Vorführung, Lagerung und Erfassung technischer Objekte.” Neue Museumskunde 24.2 (1981): 103-08. Fontenelle, Bernard Le Bovier de. Conversations on the Plurality of Worlds. Trans. H.A. Hargreaves. Berkeley and Los Angeles: University of California Press, 1990. Freundlich, Francis. Le monde du jeu à Paris 1715-1800. Paris: Albin Michel, 1995. Furetière, Antoine. Dictionnaire universel . . . 3 vols. Den Haag and Rotterdam, 1690. Gandelman, Claude. “Der Gestus des Zeigens.” Der Betrachter ist im Bild. Kunstwissenschaft und Rezeptionsästhetik. Ed. Wolfgang Kemp. Berlin: Reimer, 1992. 71-93. Gazette d’Amsterdam (February 14, 1675). Gipper, Andreas. Wunderbare Wissenschaft. Literarische Strategien naturwissenschaftlicher Vulgarisierung in Frankreich. Munich: Fink, 2002. Hallyn, Fernand. “Aspects de la problématique de l’illusion chez Descartes.” L’illusion au XVIIe siècle (= Littératures classiques, vol. 44). Ed. Patrick Dandrey and Georges Forestier. Paris: Champion, 2002. 284-304. Horn, Angelica. “Das Experiment der Zentralperspektive. Filippo Brunelleschi und René Descartes.” Descartes im Diskurs der Neuzeit. Ed. Wilhelm Friedrich Niebel and Herbert Schnädelbach. Frankfurt a.M.: Suhrkamp, 2000. 9-32. Horn-Monval, Madeleine. “La grande machinerie théâtrale et ses origines.” Revue d’histoire du théâtre 4 (1958): 291-308. Jakob, Karlheinz. Maschine, Mentales Modell, Metapher. Studien zur Semantik und Geschichte der Techniksprache. Tübingen: Niemeyer, 1991. Krafft, Fritz. Dynamische und statische Betrachtungsweise in der antiken Mechanik. Wiesbaden: Steiner, 1970. La Bruyère, Jean de. Œuvres complètes. Ed. Julien Benda. Paris: Gallimard, 1951. Lawrenson, Thomas Edward. The French Stage and the Playhouse in the XVIIth Century. A Study in the Advent of the Italian Order. New York: Ams Press, 1986. Leibniz, Gottfried Wilhelm. “Drôle de Pensée, touchant une nouvelle sorte de REPRESENTATIONS.” Sämtliche Schriften und Briefe. Ed. Akademie der Wissenschaften zu Berlin. Berlin: Akademie Verlag, 1970. Series IV, vol. I, 562-68 and notes 694-96.
The Machine as Spectacle
175
Leibniz, Gottfried Wilhelm. New Essays on Human Understanding. Ed. and trans. Peter Remnant and Jonathan Bennet. Cambridge: Cambridge University Press, 1996. Leibniz, Gottfried Wilhelm. Philosophical Texts. Ed. and trans. R.S. Woolhouse and Richard Francks. Oxford: Oxford University Press, 1998. Menestrier, Claude-François. Traité des Tournois, Ioustes, Caroussels, et autres Spectacles publics. Lyon, 1669. Mercure galant. Paris, 1673. Vol. III. Mesnadière, Jules de la. La Poëtique. Paris, 1640. Moynet, Jean. L’envers du théatre. Machines et décorations. Béziers: Hachette, 1990 [facsimile of the edition Paris, 1873]. Nicklaus, Hans Georg. “Opern des Barock als technisches Spektakel.” FrühneuzeitInfo 1 (2004): 40-46. Nobis, Heribert M. “Die wissenschaftstheoretische Bedeutung der Quaestiones Mechanicae.” Der Wissenschaftsbegriff. Historische und systematische Untersuchungen. Ed. Alwin Diemer. Meisenheim: Hain, 1970. 47-63. Parmentier, Marc, ed. L’estime des apparences. 21 manuscripts de Leibniz sur les probabilités, la théorie des jeux, l’espérance de vie. Paris: Vrin, 1995. Piemme, Jean-Marie. “L’utile dulci ou la convergence des nécessités. Recherches historiques sur les causes de l’adoption de la règle scaligérienne de l’utilité, par les théoriciens de 1630.” Revue d’histoire du théâtre 2 (1969): 118-33. Popplow, Marcus. “Die Verwendung von lat. ‘machina’ im Mittelalter und in der Frühen Neuzeit. Vom Baugerüst zu Zoncas mechanischem Bratenwender.” Technikgeschichte 30.1 (1993): 7-26. Popplow, Marcus. Neu, nützlich und erfindungsreich. Die Idealisierung von Technik in der Frühen Neuzeit. Münster: Waxmann, 1998. Popplow, Marcus. “Maschinenzeichnungen der Ingenieure der Renaissance.” Frühneuzeit-Info 1 (2004): 13-32. Pure, Michel de. Idée des spectacles anciens et nouveaux. Paris, [1668]. Rose, Paul Lawrence and Stillman Drake. “The Pseudo-Aristotelian ‘Questions of Mechanics’ in Renaissance culture.” Studies in the Renaissance 18 (1971): 65-104. Rück, Peter. “Die Dynamik mittelalterlicher Zeitmaße und die mechanische Uhr.” Die Mechanik in den Künsten. Studien zur ästhetischen Bedeutung von Naturwissenschaft und Technologie. Ed. Hanno Möbius and Jörg Jochen Berns. Marburg: Jonas Verlag, 1990. 17-30. Sabbattini, Nicola. Pratica di fabricar scene, e machine ne’ teatri . . . Ravenna, 1638. Schmidt-Biggemann, Wilhelm. “Maschine.” Historisches Wörterbuch der Philosophie. Ed. Joachim Ritter. Darmstadt: Wissenschaftliche Buchgesellschaft, 1980. Vol. 5, 790-802. Schütte, Ulrich. “Architekt und Ingenieur.” Architekt und Ingenieur. Baumeister in Krieg und Frieden. [Exhibit. cat.] Wolfenbüttel: Herzog August Bibliothek, 1984. 18-31. Stöcklein, Ansgar. Leitbilder der Technik. Biblische Tradition und technischer Fortschritt. Munich: Moos, 1969. Veranzio, Fausto. Machinae Novae Fausti Verantii siceni cum declaratione Latina, Italica, Hispanica, Gallica, et Germanica. Venice, [1600]. Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard University Press, 1962. Wiedeburg, Paul. Der junge Leibniz. Das Reich und Europa. Part II: Paris. Vol. I: Europäische Politik. Wiesbaden: Steiner, 1970. Zeising, Heinrich. Theatri machinarum [6 parts]. Leipzig, 1607-1614.
LUDGER SCHWARTE
The Anatomy of the Brain as Instrumentalization of Reason What could it mean to publicly make use of reason? Immanuel Kant’s definition of the Enlightenment1 where, under laboratory conditions, reason publicly coerces understanding into a systematic unity transforming anything sensual into the material of subjugation, has been criticized by Max Horkheimer and Theodor Adorno as the instrumentalization of reason.2 Even the idea of a system of concepts making use of understanding, language, and perception to explain itself appears to establish the autonomy of the subject.3 To what extent, however, can reason be considered an instrument at all, and become operationalized? 1. Instrument – Tool – Organ An instrument is usually understood as an object of use with which one can intervene in the structure of the environment. An instrument is then interpreted as a tool with which the user brings about a change in the form, the position, or the condition of an object. This view implies that the instrument is an object separated from the body of the user, so that a hand is no more an instrument than a snail shell. Nor, according to this notion, is it a product of the body that uses it. The advantage of this in1
2
3
Immanuel Kant. “An Answer to the Question: ‘What is Enlightenment?’” Political Writings. Ed. H.S. Reiss. Trans. H.B. Nisbet. Cambridge: Cambridge University Press, 1991. 54-59. Cf. also “What is Orientation in Thinking?” Ibid. 235-50 and ibid. 209. Cf. Max Horkheimer and Theodor W. Adorno. Dialectic of Enlightenment. Philosophical Fragments. Ed. Gunzelin Schmid Noerr. Trans. Edmund Jephcott. Stanford: Stanford University Press, 2002. 89, 94. Cf. also Theodor W. Adorno. Kant’s Critique of Pure Reason. Ed. Rolf Tiedemann. Trans. Rodney Livingstone. Cambridge: Polity Press, 2001. 62. Immanuel Kant. Prolegomena to any Future Metaphysics. Ed. Gary Hatfield. Cambridge: Cambridge University Press, 1997. § 57, 104-11.
The Anatomy of the Brain as Instrumentalization of Reason
177
terpretation is its universal applicability, inasmuch as it is characterized logically, and makes no use of psychological or social criteria (e.g. intention or profession). However, language and perception are also described as instruments. They can be practiced with the support of instruments or even independently of mere appliances. Not only specific products of the body, but also the body itself and its modification are used instrumentally. The phenomena of symbiosis, mimesis, and parasitism are linked to an instrumentalization of the world. If we do not want to reduce the concept of the instrument to that of a tool or weapon, we should also consider instruments of passivity. Instrumentalization assumes the ability to recognize possible applications. The potential user of an instrument never simply follows the rules of use. Following the rules already assumes that the user has the special mental disposition to see an object as something other. Instrumentalization does not only mean a certain makingavailable of the world, a setting up of possibilities of action, but above all a seeing-as, the intuition of alternative realities, whose foreshadowing is symbolized in the instrument. To grasp what instrumentalization does, we should not consider objects, but rather technical operations. We must understand the manipulative behavior directed towards objects, and the symbolization that makes something (an aspect, a goal) visible in relation to an object, as preparation for controlled making. Instrumentalization is the organization of operative possibilities, which can be mobilized for an anticipated goal. This organization aims at establishing material and immaterial organs – anatomical and physiological systems as well as spaces and times. Included among such an organization of operations should be those aids enabling a body, by means of performances, to transform its competences in relation to an environment, by altering its sequence or stretching its frame of operation. The social intelligence governing the use of instruments is initially defined by such an ecology of organized operations, which in the organ always also reveals the exterior, and consequently describes the type of collectivity where an operation is carried out. With this idea of instrumentalization, objects become dependent on operations, not the other way round. This change of perspective demonstrates to what extent the traditional idea of an instrument was defined by intervention or aggression. Alternatively, we could show how keeping at a distance is a more fundamental and far-reaching function of the instrument.4 Clearing ground, 4
Cf. Dominique Lestel. Les origines animales de la culture. Paris: Flammarion, 2001. 67-70.
178
Ludger Schwarte
taming fire, setting up a hearth, fixing a starting point are therefore first of all fundamentally instrumentalizing operations going to the center of the body. They are directed inward: instrumentalization is the creation of a controllable interior space. If Kant calls upon us to publicly make use of our reason, and to look for the highest touchstone of truth in ourselves, then we should enlighten ourselves about our inner space and our checkpoints, and also make this visible to others. We should draw limits to reason. We should make it into our organ. 2. Body: Model and Instrument The idea that the application of reason is tied to the corporeal has accompanied the search for the organic realization of thought since the early modern period. Even before René Descartes, the anatomy of the brain was considered the touchstone for the theoretical determining of reason’s limits. That we have a universally uniform physique, that we have nerves, a heart, a bladder, a brain, is in no way a simple finding, but rather an interpretation linked with particular operative acts and also with the functions attributed to these body parts. For Girolamo Fabricius of Aquapendente the aim of anatomy was to demonstrate not only the structure of the body, but “totius animalis fabricae theatrum.” He described the interaction of the structure based on functions which distinguished the animated as soul-edifice, fabrica animalis. Consequently, the theatrum anatomicum Fabricius had built in Padua in 1584 was a demonstration instrument where the action of an organ could be demonstrated and, in an Aristotelian way, attributed a use.5 The exhibition of functions also accompanied his investigation of the brain; here he rejected the deployment of drills and rough tools, and favored the simple use of the fourpart conical trepanning crown, which worked as silently as possible. He wrapped hammer and chisel in canvas for the protection of the patients. For fractures he used a triploid. The triploid, an upside-down tricorn, in which a pointed drill produces the fractures, invented by Hans von Gersdorff around 1530, gives a glimpse of the activity of the brain. But 5
Based on the vivisection of animals, Fabricius published books on language and its instruments (Über die Sprache und ihre Instrumente), on the eye (Über das Auge, das Organ des Sehens), and on respiration and its instruments (Über die Atmung und ihre Instrumente). Cf. Hieronymus Fabricius ab Aquapendente. Opera omnia anatomica. Leipzig, 1687.
The Anatomy of the Brain as Instrumentalization of Reason
179
what does one see when one looks at a brain freed of bone and skin? To be able to determine the structural unity of this body part or that organ, we usually assume a universal function. The notion that there are universal functions to which particular organic dispositions are attributed and which however exist to a certain extent independently of their respective material realization is already linked to a certain philosophical position and also to particular research instruments. Only both together can reveal the ‘material’ side of the functions. The construction of the universal human body can be understood from the example of Realdo Colombo. Calling on over fifteen years of dissecting experience, and distilled from his inspection of six hundred skulls in the Campus Sanctus in Rome and the Hospital S. Maria Nuova in Florence, Colombo believed in 1559 that he could represent the human body in its pure form, free of individual characteristics. This body, corresponding to the plan of nature, he named universal. Following his book on this universal body, the “universa anatome,” Colombo commented on all individual characteristics, which are correspondingly treated as monstrosities and pathologies. Vivisection experiments, provided they are carried out before a notable audience, were also considered by Colombo to be proper anatomical demonstrations. Since he was not merely (like Mondino dei Luzzi, Niccolo Massa, and Berengario da Carpi) interested in practical knowledge, his book aimed at the systematic construction of the entire body. This so-called ‘synthetic’ procedure attempted, in analogy to nature, to start from basic components. Consequently, Colombo demonstrated his anatomical knowledge by starting from the Aristotelian “principles,” the basic components of the body.6 Assuming I have a particular understanding of the brain’s activity, I can only “discover,” as the material equivalent of this understanding, what is first of all compatible with my body model. The model of the body as an instrument of cognition relates my cognitive ability (e.g. my perception) to the thing which cognizes (e.g. an organ). It organizes my range of operations. A body model is therefore an instrument determining the qualitative unity of a quantity of material. This refers particularly to the human body as the unity of perceiving and perceived. The ambiguousness of the concept of body had already been noted by Descartes.7 For Descartes the 6 7
Cf. Realdo Colombo. De re anatomice libri XV/XVI. Venice, 1559. René Descartes. “Lettre de Mr. Des Cartes au R. Père Mesland, 9 Fevrier 1645.” Œuvres complètes. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1999. Vol. 4, 166.
180
Ludger Schwarte
unity of the body could not be derived from the material components, as these increased or decreased with digestion. Only the tendency to record mental functions and let oneself be ‘informed’ by them brings a sense of structural unity.8 The soul steered the body through “esprits animaux,” which, from the brain cavities, stimulated the nerves. The nerves, for their part, moved the muscles as water in the grottos and the fountains of the royal garden moved “diverse machines” and even played instruments according to the arrangement of pipes. In Descartes’s machine allegory, the nerves are comparable with pipes, the “life spirit” however, for whose source the heart and the brain cavities would be the openings, with water.9 For these hydraulic machines, everything has a function. The heart is the inner source of movement. The brain is the fountain with which the “reasonable soul” steers the movement.10 In spite of the fundamental separation of mind and body, the brain in Descartes’s body model is attributed the central steering function. The functioning of the brain can be improved by the appropriate means, just as the ability to see can be improved by lenses. But do we think with the brain? Today, for many, this seems obvious. But it depends on the notion one has of thinking. Joseph Beuys, in any case, once said that he thought with his knee. 3. Brain Anatomy and Visualization The conviction that we think with the brain stems from experimental procedures and representational techniques of anatomy that have existed since antiquity. It is said of Alcmaeon of Croton that he dissected animals and investigated the development of chicken embryos, thus establishing that the fetuses first develop the head. In addition he severed the visual nerve of an animal and discovered a connection between the eye and the brain. Consequently, he was the first to suggest that the brain must be the seat of thought. All senses are in one way or another dependent on the brain. As soon as the brain is disturbed, the senses cannot act.11 8 9 10
11
Descartes. Œuvres complètes. Vol. 8, 315: “anima totum corpus informet.” Descartes. Œuvres complètes. Vol. 11, 130. Ibid. 131. The editors Adam and Tannery name mechanical theaters that he could have seen in Fontainbleau as a source for Descartes’s hydraulic machine paradigm. Alkmaion. “24, A 10.” Die Fragmente der Vorsokratiker. Ed. Hermann Diels and Walther Kranz. Berlin: Weidmann, 1951. 212.
The Anatomy of the Brain as Instrumentalization of Reason
181
Certain functions of the soul were also assigned to the brain by Plato and Democritus. Since the Greeks interpreted the whole activity of the soul as thought, and not only the rational operations, which they ascribed to the brain, Socrates could ask in Phaedo: And is it blood that we think with, or air, or fire? Or is it none of these, but the brain that provides the senses of hearing and seeing and smelling, from which memory and judgment come to be; and is it from memory and judgment, when they’ve acquired stability, that knowledge comes to be accordingly?12
For Aristotle, however, the brain was nothing more than a cooling device for the blood, which was larger in humans than in many other animals only because of the former’s higher body temperature. In his zoological writings Historia animalium, De generatione animalium, De partibus animalium, De motu animalium, and De incessu animalium, Aristotle established a general method for the observation of living bodies using dissection and vivisection. Dissection investigates nothing but this merging of form and function, where cutting first of all brings about the form. This inference of anatomy from physiology can be called “the anatomical method.” The anatomical method regards dissection, the exact understanding of structure, as the ideal path to the physiological discovery of function.13 The incisions in the body, the sections through the brain, would then always correspond to the idea that their structure can be described cartographically.14 The techniques of cutting as well as the hierarchy of social competences such as imagination, memory, and learning would be inscribed into these charts, corresponding to an order of knowledge. Anatomy explains how the brain is constructed, how lines and cavities are connected, and how the whole, based on these connections, functions. Nowhere is it clearer than in the field of anatomy how much has to be discarded and destroyed to make visible the functional coherence of a structure. As late as 1503, in Gregor Reisch’s encyclopaedia, an illustration of the human head appeared which, following the medieval model of Nemesios of Emesa, distributes imaginative, cognitive, and memory functions over the three brain ventricles: a fairly rough chart which reduces an intact skull to two dimensions, and from left to right outlines on this skull the ventricles lying beneath. 12 13 14
Plato. Phaedo. Trans. David Gallop. Oxford: Clarendon Press, 1975. 47-48. William F. Bynum. “The Anatomical Method, Natural Theology, and the Functions of the Brain.” Isis 64 (1973): 446. Cf. Edwin Clarke and Kenneth Dewhurst. An Ilustrated History of Brain Function. Berkeley: University of California Press, 1972. 26.
182
Ludger Schwarte
Leonardo da Vinci’s early drawing of the human brain from ca. 1490 (in Windsor Castle Library) also shows three conventional cells and their connection to optical nerves.15 Over the next fifteen years Leonardo dissected hundreds of animal and human corpses. He attempted the preparation of a brain by pouring fluid wax into the brain chambers. As soon as this had hardened, he removed the brain tissue. This made totally different structures visible. The brain acquired volume; connections dissolved into the finest branching. Leonardo’s preparation procedure suggested that this fluid wax could grasp the essence of the brain. But what should be counted as part of the brain in the first place? Here visualization plays an essential role. Galen still considered the spinal cord as part of the brain, responsible in particular for the lower parts.16 Depictions of the skull differ widely in their definition of limits. This is not only a result of a clearly different definition of what belongs to the brain, but also due to techniques of cutting and making visible, extending from anatomical instruments and processes of cutting, through language, up to the various depictions. Here, the processes of opening up and outlining, of producing and destroying overlap. The illustrations Thomas Willis used in 1664 to present his understanding of the brain were made by the architect Christopher Wren. Compared with Descartes’s illustration from 1662,17 one notices that Wren’s manner of drawing gives more emphasis to the functional architecture than the natural view, and in Cerebri Anatome, for example, draws attention, by the black and white contrast, to the recently discovered “Willis Circle” around the optical chiasma.18 Without the clarifica15
16
17 18
Cf. Clarke and Dewhurst. Illustrated History. 51, and Edwin M. Todd. The Neuroanatomy of Leonardo da Vinci. Santa Barbara: Capra Press, 1983. Gasparo Ferdinando Felice Fontana will later develop this technique further, and thus contribute to the development of the Florentine “Museo della Specola.” An observation by Thomas Willis recalls this uncertainty: “The Anatomy of the Brain.” The Remaining Medical Works of that Famous and Renowned Physician Dr. Thomas Willis. Trans. Samuel Pordage. 2 vols. Montreal, 1965 [facsimile of the London edition, 1681]. Vol. 2, 55. In his work Cerebri Anatome (1664) Willis distinguishes almost throughout between “Cerebrum,” “Medulla oblongata,” und “Cerebellum,” without developing a master term. Descartes. De Homine. Leiden, 1662. 188, Fig. LIII. Thomas Willis. Cerebri Anatome: cui accessit nervorum descriptio et usus. London, 1664. 13. A few of the drawings in Willis’s book Trevor Hughes summarized as “discoveries.” J. Trevor Hughes. Thomas Willis. London: Royal Society of Medicine Services Limited, 1991. 63. Cf. Clarke and Dewhurst. Illustrated History. 68, 71.
The Anatomy of the Brain as Instrumentalization of Reason
183
tion of a determining function, however, we only see, as Nils Stensen had already observed, a grey mass.19 The processes of visualization – this is my thesis – affect our understanding of thought and the practices of the brain. By practices of the brain I do not only mean dissection, but what is understood by brain, and what this brain, organized by the current body schema, can do. Depending on the body schema applied, the brain can appear in different roles. For example as intestines (Andreas Vesalius). As late as 1684, Raymond de Vieussens clung to the ancient notion that the convolutions of the brain and intestines are alike. In his Neurographia universalis he published an illustration of a brain whose convolutions (gyri) are intestines without any functional differentiation.20 If one takes several brains and observes them with the eyes of a technical draughtsperson, the gyri appear different. When comparing different brains Willis commented in 1664 that the surface organization of the brain was not at all chaotic. The convolutions that Willis cuts open reveal the order of a mechanism. His illustrations are highly hierarchical user instructions.21 Another example of the significance of visualization procedures is the discovery of brain fluid. Until the mid-eighteenth century, brain fluid had been overlooked simply because the anatomists had cut off the whole head for dissection so that the fluid had run out without anybody noticing.22 The representation of the brain also influenced the definition and valorization of particular mental properties. Visualization aims to match the architectonics of reason with the topography of the brain. To achieve this, the brain must, firstly, be identified as the most important organ. Secondly, the brain should offer insight into how reason is constructed. Thirdly, the body schema is given a pyramidal form instead of, for example, a spherical one. Should, fourthly, the brain no longer be hailed as the seat of the soul, but as the organ of cognition, then other mental properties are put in the shade. We should consider the brain as an object where the highest that we humans are capable of takes place. In addition, we should observe and arrange it in such a way that it enables us to check if we are thinking (correctly). Demonstration and manipula-
19 20 21 22
Niels Stensen. Sur L’anatomie du cerveau à Messieurs de l’Assemblée, qui se fait chez Monsieur Thevenot. Paris, 1669. 4. Raymond de Vieussens. Neurographia universalis. Lyon, 1684. Willis. Medical Works. 74. Domenico Felice Antonio Cotugno calls this method “ridiculous” in his De ischiade nervosa commentarius (1764).
184
Ludger Schwarte
tion overlap in experiment. Experiments are “instrumental doctrines where we learn how to progress from the known to the unknown.”23 Based on what demonstration, however, is even the skeptic Stensen, in 1669, so sure that the brain is the most important organ of the soul24 – that instrument with which it carries out such admirable things? 4. Do we Think with the Brain? With the Heart? Or with the Intestines? Until the seventeenth century there were still a few noted followers of the Aristotelian notion that the heart is the seat of the soul. There were experimental findings such as those of Francesco Redi, who removed the brain of a tortoise and reported that it lived and padded around for another year. Berengario da Carpi, Gabriele Fallopio, Ambroise Paré, and Volcher Coiter among others established that not all loss of brain substance leads to death. Coiter carried out systematic research to discover why the movement capability of birds was sustained after the severing of the head. Robert Boyle even showed that the silkworm moth, after the removal of its head, was still able to mate.25 In 1628, William Harvey demonstrated, not merely in Descartes’s sense, the mechanics of blood circulation by explaining that the heart functions like a pump. It is often overlooked that Harvey’s main interest was a return to the Aristotelian thesis of the primacy of the heart in the functional circulation of the living body.26 Unlike Aristotle, how23
24 25
26
“Experimentum est doctrina instrumentalis qua docemur progredi a noto ad ignotum.” Hieronymus Capivaccius. Opera Omnia. Ed. Johannes-Hartmannus Beyerus. Frankfurt, 1603. 1046. Stensen. L’anatomie. 4. Since Hippocrates, it had been assumed that all injuries to the brain must necessarily end in death. Galen, however, had already noticed that brain injuries could heal. Volcher Coiter. Externarum et internarum principalium humani corporis partium tabulae atque anatomicae exercitationes observationesque variae. Nuremberg, 1572. 128; Robert Boyle. The Works of the Honourable Robert Boyle. Ed. Thomas Birch. Vol. 2. London, 1772. 71. William Harvey. Exercitatio Anatomica de Motu Cordis et Sanguinis in Animalibus. Frankfurt a.M., 1628. Harvey’s Aristotelianism has already been elaborated by Pagel. Cf. Walter Pagel. William Harvey’s Biological Ideas. Basel, New York: S. Karger, 1966. 28-83 et passim. He adopted this Aristotelianism from his teacher Fabricius in Padua. The fact that he was primarily interested in the heart was due to his philosophical presumptions. Roger French notes that Harvey’s public lectures and his vivisection experiments at the College of Physicians were the key to Harvey’s discovery and led to his acceptability. Cf. Roger French. William Harvey´s Natural Philosophy. Cambridge: Cambridge University Press, 1994. 72, 87.
The Anatomy of the Brain as Instrumentalization of Reason
185
ever, Harvey was of the opinion that neither the brain nor the heart, but the blood originally feels and steers movement.27 This opinion was shared by Thomas Hobbes who deduced from this that all thought was animated by the heart. Thoughts were only reaction delays in the brain. The reactions were essentially appetite or aversion. Free will was nothing other than a delayed (through a series of consultations) reaction, so that it could be granted to animals as well as humans.28 While up to this point29 there had been an attempt to localize all physical functions in the brain, Willis responded to such findings by separating movement and feelings from thought. He assigned thought to the cerebrum, the sensorimotor system to the cerebellum. By localizing physical life in the cerebellum30 and specifying the connection between the central nervous system and the vital organs, Willis created an organic condition for the concept of unconscious thought. This was also a dramatic shift in the understanding of what the brain actually was, since, as Willis himself remarks, the ‘ancients’ did not consider the function of the cerebellum at all.31 The main experiment for the separation of the two systems involved the injury or removal of the cerebellum where, within a very short time, circulation and respiration stopped. This led to death. Sudden death after
27
28
29
30
31
A more precise comparison of Harvey and Descartes can be found in Thomas Fuchs. Die Mechanisierung des Herzens. Frankfurt a.M.: Suhrkamp, 1992. William Harvey. “An Anatomical Disquisition on the Motion of the Heart and Blood in Animals.” The Works of William Harvey. Ed. and trans. Robert Willis. London, 1847. 82, 387. Since the blood is genealogically prior to the brain, not all feeling and movement can stem from the brain. Also, as long as the brain is nothing other than ‘a limpid fluid,’ the body of the embryo will contract “if lightly pricked . . . and twist itself like a worm or caterpillar, so that it is very evidently possessed of sensation.” Ibid. 430. Cf. Harvey. “On Generation.” Ibid. 433. Thomas Hobbes. “Elements of Philosophy. The first section, concerning the body (De corpore).” The English Works. Ed. William Molesworth. London: Bohn, 1839. 392, 400, 407 et passim. Cf. Michael Kutzer. Anatomie des Wahnsinns. Geisteskrankheit im medizinischen Denken der frühen Neuzeit und die Anfänge der pathologischen Anatomie. Hürtgenwald: Guido Pressler, 1998. 82. Willis locates perception in the striate body, the imagination in the corpus callosum, the memory in the gyrus, the instinct in the midbrain, and in the cerebellum, the vital functions such as breathing, the heart function, the activities of the intestines. Willis. Medical Works. 101. Ibid. 111. The output and distribution of the perception spirits regulates the medulla, the white brain marrow. This is proved by Willis in the following an experiment: a hen, whose tiny cortex is pricked by a needle, lives and keeps walking. For Willis, therefore, the cerebral cortex as the seat of life and spirit of movement is ruled out. Willis. Cerebri Anatome. 77. Cf. ibid. 186.
186
Ludger Schwarte
lesions to the cerebellum was considered the direct proof, survival after the destruction of the cerebrum the indirect proof for his model.32 The methods and instruments used in these experiments depended on the state of surgery, particularly on the technique of trepanation. The implementation of trepanation, however, caused, in the case of many experiments on animals, misleading results since no distinction was made between the symptoms resulting from bone fragments, bleeding, etc. and symptoms of the deficiency of the sought-after systematic connection. Nor could one speak of an exact localization, since the sharp instruments – perfected since Ambroise Paré33 – were inserted through the trepan opening and poked into the depths of the brain so that the oblongata or the blood vessel were usually also injured.34 In addition, the time factor remained totally neglected, so that symptoms of initial stimuli were confused with lasting malfunctions. With the choice of test animals, attention was paid neither to the species nor to individual or age-related characteristics. Instruments, methods and the relationship of the living beings with each other corresponded to a body schema that had taken shape in relation to hydraulic machines. Willis tried, by injecting dye or ink into the main arteries of the brain, to investigate the circulation of the blood and consequently the functioning of the brain.35 He also drove nails into the skullcap to observe how the animal trembled, cramped, and then died. From the brain’s destruction by the nail, and subsequent death, however, it does not necessarily follow that the cerebellum controls vital functions, just as the possibility of alleviating the suffering of Parkinson’s patients by using a brain pacemaker does not mean that the cause of the illness is a missing electrical pulse. 32
33 34
35
Ibid. Chap. XVI, c. 24, 196. Cf. Max Neuburger. Historical Development of Experimental Brain and Spinal Cord Physiology before Flourens. Ed. and trans. Edwin Clarke. Baltimore: Johns Hopkins University Press, 1981. Cf. Ambroise Paré. Les œuvres . . . de l’anatomie que des instruments de chirurgie. Paris, 1585. Cf. the illustrations in Denis Diderot and Jean LeRond d’Alembert. Encyclopédie, ou Dictionnaire Raisonné des Sciences, des Arts et des Métiers. Paris, 1763. Plate CXII (trepanation instruments), plate XVII (representation of a trepanation using a “Bogen-Kontrepan”). Cf. Ludwig Choulant. History and Bibliography of Anatomic Illustration. Trans. Mortimer Frank. New York: Hafner, 1945. 228. Willis, on the other hand, initiated a kind of anatomical experiment where, using rough methods, individual parts of the brain were injured to reach an “increasingly precise assignation of single moments of the ‘mental faculty’ to substructures of this cerebral organ.” Olaf Breidbach. Die Materialisierung des Ichs. Zur Geschichte der Hirnforschung im 19. und 20. Jahrhundert. Frankfurt a.M.: Suhrkamp, 1997. 58.
The Anatomy of the Brain as Instrumentalization of Reason
187
As long as a living organism can be understood as a machine whose functions are derived from structural characteristics, the functioning of a machine is the frame of reference for the body schema. For Willis’s anatomical method, however, structureless functions and functionless structures are dead ends that put the whole body schema in question. Thought can be one such dead end. This is why Willis assumed on the one hand that the complexity of the gyri and the intelligence were direct correlates.36 The human nervous system was more refined because more complicated than that of other living beings. Here he saw anatomical proof for placing man as rational being at the summit of creation. On the other hand, however, he observed that the nervous systems of humans and animals hardly differed. The intimate link between brain structure and the function of the thinking organ can be seen in Willis’s case in an exemplary way where at one point he compares the brain with a machine: “By the effects, and by comparing rationally the Faculties, and Acts, with the workmanship of the Machine, we may at least conjecture, what sort of works of the Animal Function, are performed in these or those, or within some other parts of the head.”37 The purpose of the structural elements lies in the creation of thought-acts. The inference of action from the plan and its mechanical implementation guides Willis’s anatomy of the cerebral cortex. On the one hand, this approach completely ignores the metabolic processes in the brain. On the other hand, Willis is convinced that the cerebrum must be the seat of certain unusual abilities of humans merely because of the observation that this part of the brain appears to be constructed in a more complex way. Here, the assumption of a brain architecture (“fabrica cerebri”) is much more than an analogy. For Willis, the mental abilities of all living beings depend on the folds, canals and meeting points in the brain, just as the brain, because of its complex organization and heavy fortification, controls, like a capital city, the provinces of the body.38 Inscribed in Willis’s anatomical cartography is a specific early modern schema of the body politic. Only this Body Politick makes the correla36 37
38
Willis. Cerebri Anatome. 65. Cf. Kenneth Dewhurst. Willis’s Oxford Lectures. Oxford: Sandford Publications, 1980. 55. Thomas Willis. “Two Discourses concerning the Soul of Brutes.” Dr. Willis’s Practice of Physick. Ed. Samuel Pordage. London, 1684. 27. In other places Willis distinguishes the instinct-like efforts, which the cerebrum controls, from the ability to learn, which might cultivate musicality, judgment, and intelligence. Cf. Dewhurst. Willis’s Oxford Lectures. 147. Therefore Willis calls the skull a “Castellum,” the corpus callosum a “publicum emporium.” Willis. Cerebrium Anatome. 122, 129.
188
Ludger Schwarte
tion between freedom, learning and control and the elaborate construction (diversity, differentiation, and linking) of the human brain plausible. The same cultural pattern underlies Willis’s remark that the structure of the cerebellum hardly varies from species to species, as he assumes that the vital functions of all living beings are the same.39 The structural differences between various species are quantitative;40 the construction of the brain is hardly distinguished by shape, but rather by variations of scale.41 It is particularly the similarity of the “figure and size” of human and ape brains which leads him to the conclusion that no brain structure can be distinguished which can be considered as uniquely human.42 This leads him to the conviction adopted from Pierre Gassendi that man is “a two-soul’d Animal”43 and possesses an animalistic-sensitive as well as a rational, exclusively human soul. Although he rejected the Cartesian localization of the soul in the pineal gland,44 “because Animals, which seem to be almost quite destitute of Imagination, Memory, and other superior Powers of the Soul, have this glandula or kernal large and fair enough,”45 Willis still believed he could attribute certain functions of the rational soul to the corpus callosum. According to Willis, the corpus callosum was the region of the brain where the main abilities of knowledge and understanding of the soul reside.46 In particular, the imagination was organized there. “The Rational Soul . . . depends very much, as to its operation, on the Fantasy, without the help of which, it can know or under39
40
41 42 43 44
45 46
Willis. Medical Works. 91f. Cf. Alfred Meyer and Raymond Hierons. “On Thomas Willis’s Concepts of Neurophysiology.” Medical History 9 (1965): 1-15, 14255. Alfred Meyer. Historical Aspects of Cerebral Anatomy. New York and London: Oxford University Press, 1971. Hughes. Thomas Willis. 62. The certainty of his method was not the result but also the motivation for an extensive dissection praxis. He boasted that he had “slain so many Victims, whole Hecatombs almost of all Animals, in the Anatomical Court.” Willis. Medical Works. “Dedication.” 100. Willis. Practice of Physick. 44. Cf. Willis. Medical Works. 132; Willis. Cerebri Anatome. 124. Willis. “Preface.” Practice of Physick. The body is the organ of a material soul. Ibid. 7. Descartes assumes that the soul is spread out through the whole body. Cf. Descartes. “Les Passions de l’Ame (1649).” Œuvres complètes. Vol. 11, 351. Descartes argues, concerning the attribution of the heart as seat of the soul, that not the heart but the nerves connected to it are responsible for the location of pain. The soul is the feeling not the felt. If that is the case, nothing speaks for the location of thought in the brain. Willis. Medical Works. 87. Willis. Practice of Physick. 153.
The Anatomy of the Brain as Instrumentalization of Reason
189
stand nothing.”47 The imagination is then separated from rationality and assigned to the brain. Although people also vary in their intelligence and their powers of judgment, Willis, like Descartes,48 clung to the notion that everybody is given a rational soul in the same measure. The ability to act depends merely on the construction of the sensitive soul.49 That the construction of the brain might be decisive for the application of rational capabilities, Willis inferred from physical defects. Should the areas of the brain where the imagination and memory are situated be defective, then an absent power of judgment would follow, “stupidity and foolishness.”50 The mental functions of the rational soul could therefore be localized in the brain, while strong emotions such as desire, anger, fear, and grief exercise, in Willis’s model, had a direct influence on the heart. The force of these reactions and sensations was sensed in the thorax rather than in the head. He gave his reasons using, once again, an anatomical comparison: whereas animals draw almost all their cardiac enervation from the vagus nerves, man has additional nerve fibers. Willis thus concluded that the twofold enervation of the human heart, in contrast to animals, demonstrated the close link between the heart and the brain, and consequently the higher disposition of man for moral acts.51 As a result, Willis derived the internal architecture of the rational soul from the anatomy of the brain and the nervous system. It could therefore be seriously handicapped by physical defects, but also cured using physical methods. 47 48
49 50 51
Ibid. 41. Descartes poses this question in Discours de la Méthode. He conjectures that the ability is given to all men in the same measure, but that it develops itself fully in none. In connection with brain anatomy, the separation between soul and mind which starts to emerge in this period can also be seen. Therefore Johann Franz Budde writes, following the physiologist Friedrich Hoffmann, that the body/soul philosophies are divided into two camps: The first distinguish in man two parts, namely body and soul; the other however three, namely body, soul, and mind. The mind (mens) is here the movement principle of the body machine, the soul (anima) that which distinguishes animal from man and which in the body functions through the blood, or the spiritus animales. Cf. Johann Franz Budde. Elementa philosophiae instrumentalis et theoreticae. Vol. 2. Halle, 1703. 106-08; Friedrich Hoffmann. Fundamenta Medicinae ex principiis naturae mechanicis in usu Philiatrorum succinte proposita. Halle, 1695. 15. Willis. Practice of Physick. 41. Ibid. 209-11. Willis. Medical Works. 132. Though not only morality but rationality as a whole depends on this particular nerve connection. Willis proves this by the dissection of a “fool.” Ibid. 132.
190
Ludger Schwarte
In De Motu Musculari,52 Willis discovered a power of movement in the muscles independent of brain and nerves.53 A large part of what was previously considered deliberate movement, for Willis became a simple reflex, which even single limbs were capable of. The rational soul shrank to a few functions independent of the body. Anatomy established the idea of an organic life to which the physical was now reduced. However, just as it found no structural correspondence for human thought functions,54 anatomy also discovered organs without function. The larynx of the ape may be available, though it does not use its voice organ. Perrault had already concluded from this that actors do not “perform” this or that action only because they have the appropriate organ.55 The experimental set-ups therefore concentrated harder on forcing the organs into particular reactions. From Stensen’s grey mass, a twitching tangle of nerves emerged. 52
53
54
55
Willis. “De Motu Musculari.” Opera Omnia. Vol. 1. Lyon, 1681. 674. The term “neurology,” which he used for the first time in 1664, is derived from Willis. According to Georges Canguilhem the concept of reflex movement is derived from Willis rather than Descartes. Georges Canguilhem. La formation du concept de réflexe aux XVIIe et XVIIIe siècles. Paris: P.U.F, 1955. 57, 169 et passim. Willis developed this concept in 1670 in De Motu Musculari. According to Isler, other modern terms such as animal psychology, comparative anatomy, and hormones also derive from Willis’s work. Cf. Hansruedi Isler. Thomas Willis. Ein Wegbereiter der modernen Medizin. Stuttgart: Wissenschaftliche Verlagsgesellschaft, 1965. 88-89. Cf. also Francis Joseph Cole. A History of Comparative Anatomy. From Aristotle to the Eighteenth Century. London: Macmillan, 1944. The movement impulses of the muscles surpass those of the nerves. They are set in motion by a kind of explosion which is triggered by the brain and conveyed by the nerves. Experiments with gunpowder influence this model. This attentiveness to the local sources of the motor-forces inspired Claude Perrault’s mechanical explanation, based on “ressorts” and “inhibitions.” Cf. Claude Perrault. Essais de physique. Vol. 3. Paris, 1688. 80. Cf. Mirko Grmek. La première révolution biologique, Réflexions sur la physiologie et la médecine du XVIIe siècle. Paris: Payot, 1990. 179-83. Nevertheless, two opposing discoveries stood in the way of the integration of the anatomical method into the classical notion of natural philosophy that a divine plan can also be found in the details of the structure of natural organs: the brain of man resembled in all decisive points that of the ape, but was suited to completely different functions. If Willis wants to formulate this dualism in a consistent way and assert the physiological disjunction between man and animal, then he should not aim to find one nerve to serve as the medium of morality. If the anatomical method shows the similarity of human and animal brains, then the conclusion only remains that either the supposed functional differences do not exist – and then animals are also able to speak and think, but at a different level – or that the anatomical method is not sufficient to recognize the decisive differences. Claude Perrault. Mémoires pour servir à l’histoire naturelle des animaux. Paris, 1671. 163.
The Anatomy of the Brain as Instrumentalization of Reason
191
Nevertheless, Albrecht von Haller, despite various incisions, the infliction of burns, or the application of acids, could not discern any sensitivity to pain in the brain.56 The brain could not be stimulated; it, unlike the nerves, did not move. Stimulations and sensitivity to pain seemed, however, more than thought, to indicate the control of vital functions. For Haller, the body was an organized structure of forces and reaction mechanisms. Using the example of electrical stimulus, he showed, on the one hand, that a nerve can also be stimulated when its connection to the brain is severed, and on the other hand, that “the stimulation does not, as generally assumed, arise from the nerves, but follows from the structure of the stimulated part itself.”57 From the hydraulic machine an electrical machine emerged which could be stimulated externally at every point. The electric body schema, however, is challenged by the chemical one. In his treatise Über das Organ der Seele (1796), Samuel Thomas Soemmering describes the external morphology of the human gyri and brain nerves, and supports the thesis that not the brain tissue, but the moisture of the brain cavities is the organ of the soul. Although this draws on the notion of a life force, he interprets its reactions mechanistically. In response to Soemmering’s invitation, Kant commented on the compromise of considering the organ of the soul as an ethereal-material machine.58 According to Kant, the mechanical idea of propulsion and repulsion must be expanded to include the procedures of the soul-organ. Kant demanded that at least chemical chains of action must also be included in order to produce a “dynamic organization.” The expression “seat of the soul” should already put philosophers on their guard, since the soul cannot have a place but only a virtual present which is part of the understanding. If most of us have the impression that we think with our heads, then this is actually an error in which a judgment about the cause of a sensation in the brain is confused with the sensation of the cause in the brain. Soemmering did not distinguish between the seat and the organ of the soul. The brain fluid could not be 56 57
58
Cf. Albrecht von Haller. Elementa Physiologiae corporis humani. Lausannes, 1757-66. Vol. 4, 515. Quoted in Gerhard Rudolph. “Hallers Lehre von der Irritabilität und Sensibilität.” Von Boerhaave bis Berger. Die Entwicklung der kontinentalen Physiologie im 18. und 19. Jahrhundert. Ed. Karl Eduard Rothschuh. Stuttgart: G. Fischer, 1964. 26. Cf. Michael Hagner. “Das Ende vom Seelenorgan: Über einige Beziehungen von Philosophie und Anatomie im frühen 19. Jahrhundert.” Das Gehirn – Organ der Seele? Zur Ideengeschichte der Neurobiologie. Ed. Ernst Florey and Olaf Breidbach. Berlin: Berlin Verlag, 1993. 7.
192
Ludger Schwarte
the organ of the soul “since the water, as fluid, cannot justifiably be thought of as organized, but on the contrary as being entirely without organization . . . no material is apt to constitute the immediate organ of the soul, – that good discovery has not yet reached its goal.”59 For Kant, the functional organization of the soul was not revealed spatially, but temporally, as a unity of inner contemplation. He stressed that the soul was immaterial, and not a substance.60 For reasons of method it could not be comprehended by the medical faculty, but only by the philosophical. Kant’s schematization of reason also depended on a body model that would not be thinkable without certain anatomical, physiological, and pathological discoveries. Since, according to Kant, the application of reason is not only restricted by a confusion of the senses or physical disability, but also by emotional disorders. These could be caused by the “physical confusion of the soul organs”61 or experiments with drugs.62 Kant concluded in his Versuch über die Krankheiten des Kopfes that these actually appeared in the mind [Gemüte], though the roots of these illnesses actually lay in the body, and indeed “more in the digestion parts than in the brain.”63 Healthy diet, healthy spirit [Geist]. In the Träumen eines Geistersehers, Kant criticized the Cartesians, who considered the seat of the soul in the brain “like a spider at the center of its web,” which is knocked and shaken by the nerves on the one hand, but also moves “the ropes and levers of the whole machine.” This he counters with the belief: “where I feel, there I am.”64 For Kant sensation was the whole body, but reason a faculty making use of chemical processes for its realization as temporal organization of the body. Two changes stand out here: 1) for Kant, only the “rational soul” still plays a role, and 2) rationality is in principal independent of its “medium,” namely to be grasped as system. Kant enthrones philosophy as a system for “rational cognition according to concepts.” His belief in reason is a matter of the dynamic domination of the body using concepts.65 59
60 61 62 63 64 65
Immanuel Kant. Werkausgabe. 12 vols. Ed. Wilhelm Weischedel. Frankfurt a.M.: Suhrkamp, 1964. Vol. 11, 257. Cf. Samuel Thomas Soemmering. Über das Organ der Seele. Amsterdam, 1966 [facsimile of the Königsberg edition, 1796]. Cf. Immanuel Kant. Critique of Pure Reason. Trans. Norman Kemp Smith. New York: St. Martin’s Press, 1965. 76-82, 328-33, 549-73. Kant. Werkausgabe. Vol. 12, 529. Ibid. 532. Kant. Werkausgabe. Vol. 2, 900. Ibid. 932-33. Cf. Immanuel Kant. Critique of Judgment. Trans. James Creed Meredith. Oxford: Clarendon Press, 1989. 17. Kant. Werkausgabe. Vol. 10, 9.
The Anatomy of the Brain as Instrumentalization of Reason
193
Clearly particular philosophical precedents typical for the times determine the field of research. Should I look for the soul in the brain? What do I understand by thought? Does the I materialize itself, as Olaf Breidbach asks, in the brain?66 Do brains communicate, as Wolf Singer suggests,67 or people? Am I only my thought? Does thought materialize in the brain? Or, as Wittgenstein says, on a piece of paper? 5. Instrumental Practices of the Brain To finish, I will try to sketch the cultural consequences produced by alternative ideas of the organ of thought (brain/heart/the whole body). a) First of all it is important to stress that various models of research give rise to particular therapies. In the eighteenth century numerous patients were trepanned to prevent possible brain diseases.68 Trepanation [Scheintrepanationen, “Narrenschneiden”] as psychiatric treatment was not replaced until the emergence of electroshock therapy. Without the electroencephalogram, the concept of brain death would be unthinkable.69 b) Current procedures of visualization, from EEG to computer tomography, try to demonstrate an increasingly specific correlation between functions and brain regions, but tend in their interpretations to go beyond merely establishing correlations. They substantiate thought in the brain, exactly as if one can demonstrate that speaking is inherent in the tongue only because without the tongue speaking is not possible, and it becomes active while speaking. Since we are in a position to measure particular electric action potentials (“spikes”), we assume the brain processes information in the form of discharges. Though here it does not 66
67 68
69
Rorty’s answer is not very convincing: “Just as the brain is not something that ‘has’ such synapses, but is simply the agglomeration of them, so the self is not something which ‘has’ the beliefs and desires but is simply the network of such beliefs and desires.” Richard Rorty. “Non-reductive Physicalism.” Objectivity, Relativism, and Truth. Philosophical Papers. Vol. 1. Cambridge: Cambridge University Press, 1990. 123. Wolf Singer. Ein neues Menschenbild? Gespräche über Hirnforschung. Frankfurt a.M.: Suhrkamp, 2003. 56. Working in the Hôtel Dieu in Paris, Jean Méry is supposed to have lost all the patients he trepanned. Cf. Christa Habrich. Loch im Kopf. Zur Geschichte der Schädeltrepanation (= Sammelblätter des Deutschen Medizinhistorischen Museums). Ingolstadt, 2002. 9. Cf. Claudia Wiesemann. “Instrumentalisierte Instrumente: EEG, zerebrale Angiographie und die Etablierung des Hirntodkonzeptes.” Instrument – Experiment. Historische Studien. Ed. Christoph Meinel. Berlin: GNT-Verlag, 2000. 225-35.
194
Ludger Schwarte
even seem clear what a ‘neuronal code’ is – is it the discharge rate, the temporal pattern or the appearance of a “spike”? Are stimuli from the environment represented by groups of neurones (“population codes”)? Or does it depend on the chemical message material, the neurotransmitters or neuropeptides?70 What role do the glial cells which in the human brain appear much more frequently than nerve cells play here? Clearly in the current models the character of instrument is transferred to the brain. Thought activity, a cultural, a political concept, is determined by the measurability of correlative processes. c) A direct consequence of this approach is schematization of the body, which is used to judge mental phenomena. Therefore, it is not false to count Willis among the precursors of phrenology and eugenics as he was convinced that stupidity was the result of the wrong size and shape of the brain as well as an “inferior clay.”71 d) Among the cultural consequences, not least significant is the political dimension of the experimentally supported body schema. Willis described the cerebellum as a free city (“Civiati librae & municipali”),72 where life forces, independent of a government and their decisions, move independently like machines,73 if peace is not destroyed by violent passions in the forum of the brain.74 While prioritizing the heart suggests a hierarchy which puts the impulses, the life-giving and movement-guiding forces as well as the passions center stage, prioritizing the brain calls for another form of government specializing in the orchestration of movement on the basis of the administration and interpretation of the (visually) perceived.75 The instrumentalization of thought and the concurrent schematization of the body, and the adaptation of thought to the supposed structure of its organ, does not only serve as the legitimization of a particular form of control, but also demands the estimation of philosophy. This finds its clearest expression in the philosophical program of Denis Diderot, who might act as the prompter of Kant’s concept of enlighten70 71 72 73 74 75
Cf. Gerhard Roth. Das Gehirn und seine Wirklichkeit. Frankfurt a.M.: Suhrkamp, 1994. 18. Dewhurst. Willis’s Oxford Lectures. 134, 136. Willis. Cerebrium Anatome.184. Cf. Willis. Medical Works. 111. Willis. Cerebrium Anatome.198. An example might be the conception of the brain as conductor or choirmaster (“cerebrum tanquam mester del choro”). William Harvey. De Motu Locali Animalium. Ed. and trans. Gweneth Whitteridge. Cambridge: Cambridge University Press, 1959. 110.
The Anatomy of the Brain as Instrumentalization of Reason
195
ment. Diderot justified the undertaking of the encyclopaedia as the instrumentalization of the life-world with words: The right approach to philosophy, both in the past and now, would have been to apply the understanding to what has been understood; to apply understanding and experimentation to the senses; the senses to nature, and nature to the investigation of the instruments to be used; and finally, to employ these instruments for researching into and perfecting the arts, which should be set before the public to teach people to respect philosophy.76
The belief in subjective reason, though conceived by Descartes as a way of bridging the gap between science and theology, instrumentalizes thought77 tied to a social hierarchy, according to Horkheimer. By instrumentalization, Horkheimer understood the process of subjugation to subjective-social aims. Thoughts become things, machines in the apparatus of production, sense is basically considered conformist, bound to human aims.78 On the other hand, things lose all intrinsic value “if true judgments on objects, and therewith the concept of the object itself, rests solely on ‘effects’ upon the subject’s action, it is hard to understand what meaning could still be attributed to the concept ‘object.’”79 The pragmatists want, according to Horkheimer, to remodel all philosophy as experimental physics; their pride is “to think of everything just as everything is thought of in the laboratory, that is, as a question of experimentation.”80 Pragmatists attempted to bind human practice to the results of experiments, making however the value of these results dependent “on the accurate definition of ‘all the conceivable experimental phenomena,’” so that Horkheimer asks skeptically: “How is it possible to subject experimentation to the criterion of ‘being conceivable,’ if any concept – that is to say, whatever might be conceivable – depends essentially on experimentation?”81 The function of pragmatic philosophy is a social one. Only the experiment should still count. 76
77
78 79 80 81
Denis Diderot. Thoughts on the Interpretation of Nature and other Philosophical Works. Intro. David Adams. Trans. Lorna Sandler. Manchester: Clinamen Press, 1999. 43. Max Horkheimer. Critique of Instrumental Reason. Trans. Matthew J. O’Connell et al. New York: Continuum, 1974. vii. Already Friedrich Nietzsche had recognized that since Descartes’s philosophy, “reason is only an instrument.” Friedrich Nietzsche. Beyond Good and Evil. Trans. R.J. Hollingdale. Intro. Michael Tanner. London: Penguin, 1973. 114. Max Horkheimer. Eclipse of Reason. New York: Continuum, 1974. 21-22. Ibid. 45. Charles S. Peirce. Collected Papers. Cambridge: Harvard University Press, 1934. Vol. 5, 272. Horkheimer. Eclipse. 48.
196
Ludger Schwarte
The process that tends to replace the various theoretical ways to objective truth with the powerful machinery of organized research is sanctioned by philosophy, or rather is being identified with philosophy. All things in nature become identical with the phenomena they present when submitted to the practices of our laboratories.82
Here, the ‘busy experimentation’ only produces concrete answers to questions corresponding to particular social interests, a subjective, economic reason. With the reduction of reason to an instrument, thought is no longer measured by thought, but by criteria of production and social effectiveness. This is due, though, to particular practices – not considered by Horkheimer – of the brain. The anatomy of the brain justifies social subordination. Willis’s political schema for the description of the brain functions reappears in the social organization which Horkheimer diagnoses: “The leading functions of production – commanding, planning, organizing – were contrasted as pure intellect to the manual functions of production as lower, impurer form of work, the labor of slaves.” By the coupling of scientific and industrial practice, reason is both disembodied and excessive as a system, though it also becomes blind: “The neutralization of reason . . . transforms it . . . into a mere dull apparatus for registering facts. Subjective reason loses all spontaneity, productivity, power to discover and assert new kinds of content.”83 The brain is no longer an object which could contain an unsystematic truth.84 e) Finally, among the philosophical and political consequences of the instrumentalization of reason by anatomy is the discourse on the cultivation of brains for the improvement of mankind, which emerges with Juan Huarte de San Juan in 1575.85 In 1686 Bernard Le Boyer de Fontenelle wrote in his Traité de la Liberté that thought is produced by the material disposition of the brain, and that the things grasped in thoughts also leave their traces there. If this disposition constantly provides the impetus for the evaluation of a matter and therefore for virtuous or bad thoughts, Fontenelle might have been able to sketch a mechanics of the decisions guiding human practice. As a result thoughts would never be free. Fontenelle substantiates this assertion with the example of the insane, children, and dreamers whose thoughts depend equally on a psycho-physical mechanism: “What the soul wants is necessarily determined by the brain, and the will in turn excites a move82 83 84 85
Ibid. 49. Ibid. 54f. Singer. Menschenbild. 56-59. Juan Huarte de San Juan. Examen de ingenios para las ciencias. Baeza, 1575.
The Anatomy of the Brain as Instrumentalization of Reason
197
ment in the brain by which it is executed.” Because, according to Fontenelle’s conception, the thoughts leave behind traces in the physiology, he developed a targeted disciplining of the brain, since “one can install certain dispositions in the brain which determine the degree of virtuousness.”86 A coherent system of laws, punishments and rewards should socialize the brain. This discourse leads to an idolization of the brain under laboratory conditions. This can still be heard in Gilles Deleuze and Félix Guattari’s prophecy of the end of humanity: “It is the brain that thinks and not man,” as they say here, “the latter being only a cerebral crystallization. We will speak of the brain as Cézanne spoke of the landscape: man absent from, but completely within the brain.”87 Perhaps we should think with the intestines. And digest with the brain.
Translation: Benjamin Carter WORKS CITED Adorno, Theodor W. Kant’s Critique of Pure Reason. Ed. Rolf Tiedemann. Trans. Rodney Livingstone. Cambridge: Polity Press, 2001. Alkmaion. “24, A 10.” Die Fragmente der Vorsokratiker. Ed. Hermann Diels and Walther Kranz. Berlin: Weidmann, 1951. Boyle, Robert. The Works of the Honourable Robert Boyle. Ed. Thomas Birch. Vol. 2. London, 1772. Breidbach, Olaf. Die Materialisierung des Ichs. Zur Geschichte der Hirnforschung im 19. und 20. Jahrhundert. Frankfurt a.M.: Suhrkamp, 1997. Budde, Johann Franz. Elementa philosophiae instrumentalis et theoreticae. Vol. 2. Halle, 1703. Bynum, William F. “The Anatomical Method, Natural Theology, and the Functions of the Brain.” Isis 64 (1973): 445-468. Canguilhem, Georges. La formation du concept de réflexe aux XVIIe et XVIIIe siècles. Paris: P.U.F, 1955. 86
87
Bernard Le Boyer de Fontenelle. “Traité de la Liberté.” Textes choisis. Ed. Maurice Roelens. Paris: Editions sociales, 1967. 153. The Traité is also printed in Fontenelle und die Aufklärung. Ed. Werner Krauss. Munich: Fink, 1969. 283-93. Cf. also Carmelo Romeo. “Matérialisme et Déterminisme dans le Traité de la Liberté de Fontenelle.” Le Matérialisme du XVIIIe siècle et la littérature clandestine. Ed. Olivier Bloch. Paris: Vrin, 1982. 4-8. Particularly Diderot is influenced by this text. Cf. his “Lettre à M. Landois” (1756). Diderot. “Lettre à Landois du 29 juin 1756.” Correspondance. Vol. 1. Paris: Éditions de Minuit, 1955. 209. Gilles Deleuze and Félix Guattari. What is Philosophy? New York: Columbia University Press, 1994. 210.
198
Ludger Schwarte
Capivaccius, Hieronymus. Opera Omnia. Ed. Johannes-Hartmannus Beyerus. Frankfurt, 1603. Choulant, Ludwig. History and Bibliography of Anatomic Illustration. Trans. Mortimer Frank. New York: Hafner, 1945 Clarke, Edwin and Kenneth Dewhurst. An Ilustrated History of Brain Function. Berkeley: University of California Press, 1972. Coiter, Volcher. Externarum et internarum principalium humani corporis partium tabulae atque anatomicae exercitationes observationesque variae. Nuremberg, 1572. Cole, Francis Joseph. A History of Comparative Anatomy. From Aristotle to the Eighteenth Century. London: Macmillan, 1944. Colombo, Realdo. De re anatomice libri XV/XVI. Venice, 1559. Deleuze, Gilles and Félix Guattari. What is Philosophy? New York: Columbia University Press, 1994. Descartes, René. De Homine. Leiden, 1662. Descartes, René. Œuvres complètes. Ed. Ch. Adam and P. Tannery. Paris: Vrin, 1999. Dewhurst, Kenneth. Willis’s Oxford Lectures. Oxford: Sandford Publications, 1980. Diderot, Denis and Jean LeRond d’Alembert. Encyclopédie, ou Dictionnaire Raisonné des Sciences, des Arts et des Métiers. 35 vols. Ed. idem. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 1751-1780. Diderot, Denis. “Lettre à Landois du 29 juin 1756.” Correspondance. Paris: Éditions de Minuit, 1955. Vol. 1, 209. Diderot, Denis. Thoughts on the Interpretation of Nature and other Philosophical Works. Intro. David Adams. Trans. Lorna Sandler. Manchester: Clinamen Press, 1999. Fabricius ab Aquapendente, Hieronymus. Opera omnia anatomica. Leipzig, 1687. Fontenelle, Bernard Le Boyer de. “Traité de la Liberté.” Textes choisis. Ed. Maurice Roelens. Paris: Editions sociales, 1967. French, Roger. William Harvey´s Natural Philosophy. Cambridge: Cambridge University Press, 1994. Fuchs, Thomas. Die Mechanisierung des Herzens. Frankfurt a.M.: Suhrkamp, 1992. Grmek, Mirko. La première révolution biologique, Réflexions sur la physiologie et la médecine du XVIIe siècle. Paris: Payot, 1990. Habrich, Christa. Loch im Kopf. Zur Geschichte der Schädeltrepanation (= Sammelblätter des Deutschen Medizinhistorischen Museums). Ingolstadt, 2002. Hagner, Michael. “Das Ende vom Seelenorgan: Über einige Beziehungen von Philosophie und Anatomie im frühen 19. Jahrhundert.” Das Gehirn – Organ der Seele? Zur Ideengeschichte der Neurobiologie. Ed. Ernst Florey and Olaf Breidbach. Berlin: Berlin Verlag, 1993. Haller, Albrecht von. Elementa Physiologiae corporis humani. Vol. 4. Lausannes, 1757-66. Harvey, William. Exercitatio Anatomica de Motu Cordis et Sanguinis in Animalibus. Frankfurt a.M., 1628. Harvey. William. The Works of William Harvey. Ed. and trans. Robert Willis. London: Sydenham Society, 1847. Harvey, William. De Motu Locali Animalium. Ed. and trans. Gweneth Whitteridge. Cambridge: Cambridge University Press, 1959. Hobbes, Thomas. “Elements of Philosophy. The first section, concerning the body (De corpore).” The English Works. Ed. William Molesworth. London: Bohn, 1839. Hoffmann, Friedrich. Fundamenta Medicinae ex principiis naturae mechanicis in usu Philiatrorum succinte proposita. Halle, 1695. Horkheimer, Max. Critique of Instrumental Reason. Trans. Matthew J. O’Connell et al. New York: Continuum, 1974.
The Anatomy of the Brain as Instrumentalization of Reason
199
Horkheimer, Max. Eclipse of Reason. New York: Continuum, 1974. Horkheimer, Max and Theodor W. Adorno. Dialectic of Enlightenment. Philosophical Fragments. Ed. Gunzelin Schmid Noerr. Trans. Edmund Jephcott. Stanford: Stanford University Press, 2002. Huarte de San Juan, Juan. Examen de ingenios para las ciencias. Baeza, 1575. Hughes, J. Trevor. Thomas Willis. London: Royal Society of Medicine Services Limited, 1991. Isler, Hansruedi. Thomas Willis. Ein Wegbereiter der modernen Medizin. Stuttgart: Wissenschaftliche Verlagsgesellschaft, 1965. Kant, Immanuel. Werkausgabe. 12 vols. Ed. Wilhelm Weischedel. Frankfurt a.M.: Suhrkamp, 1964. Kant, Immanuel. Critique of Pure Reason. Trans. Norman Kemp Smith. New York: St. Martin’s Press, 1965. Kant, Immanuel. Critique of Judgment. Trans. James Creed Meredith. Oxford: Clarendon Press, 1989. Kant, Immanuel. Political Writings. Ed. H.S. Reiss. Trans. H.B. Nisbet. Cambridge: Cambridge University Press, 1991. Kant, Immanuel. Prolegomena to any Future Metaphysics. Ed. Gary Hatfield. Cambridge: Cambridge University Press, 1997. Krauss, Werner, ed. Fontenelle und die Aufklärung. Munich: Fink, 1969 Kutzer, Michael. Anatomie des Wahnsinns. Geisteskrankheit im medizinischen Denken der frühen Neuzeit und die Anfänge der pathologischen Anatomie. Hürtgenwald: Guido Pressler, 1998. Lestel, Dominique. Les origines animales de la culture. Paris: Flammarion, 2001. Meyer, Alfred. Historical Aspects of Cerebral Anatomy. New York and London: Oxford University Press, 1971. Meyer, Alfred and Raymond Hierons. “On Thomas Willis’s Concepts of Neurophysiology.” Medical History 9 (1965): 1-15, 142-155. Neuburger, Max. Historical Development of Experimental Brain and Spinal Cord Physiology before Flourens. Ed. and trans. Edwin Clarke. Baltimore: Johns Hopkins University Press, 1981. Nietzsche, Friedrich. Beyond Good and Evil. Trans. R.J. Hollingdale. Intro. Michael Tanner. London: Penguin, 1973. Pagel, Walter. William Harvey’s Biological Ideas. Basel, New York: S. Karger, 1966. Paré, Ambroise. Les œuvres . . . de l’anatomie que des instruments de chirurgie. Paris, 1585. Perrault, Claude. Essais de physique. Vol. 3. Paris, 1688. Perrault, Claude. Mémoires pour servir à l’histoire naturelle des animaux. Paris, 1671. Peirce, Charles S. Collected Papers. Vol. 5. Cambridge: Harvard University Press, 1934. Plato. Phaedo. Trans. David Gallop. Oxford: Clarendon Press, 1975. Romeo, Carmelo. “Matérialisme et Déterminisme dans le Traité de la Liberté de Fontenelle.” Le Matérialisme du XVIIIe siècle et la littérature clandestine. Ed. Olivier Bloch. Paris: Vrin, 1982. Rorty, Richard. “Non-reductive Physicalism.” Objectivity, Relativism, and Truth. Philosophical Papers. Vol. 1. Cambridge: Cambridge University Press, 1990. Roth, Gerhard. Das Gehirn und seine Wirklichkeit. Frankfurt a.M.: Suhrkamp, 1994. Rudolph, Gerhard. “Hallers Lehre von der Irritabilität und Sensibilität.” Von Boerhaave bis Berger. Die Entwicklung der kontinentalen Physiologie im 18. und 19. Jahrhundert. Ed. Karl Eduard Rothschuh. Stuttgart: G. Fischer, 1964. Singer, Wolf. Ein neues Menschenbild? Gespräche über Hirnforschung. Frankfurt a.M.: Suhrkamp, 2003.
200
Ludger Schwarte
Soemmering, Samuel Thomas. Über das Organ der Seele. Amsterdam, 1966 [facsimile of the Königsberg edition, 1796]. Stensen, Niels. Sur L’anatomie du cerveau à Messieurs de l’Assemblée, qui se fait chez Monsieur Thevenot. Paris, 1669. Todd, Edwin M. The Neuroanatomy of Leonardo da Vinci. Santa Barbara: Capra Press, 1983. Vieussens, Raymond de. Neurographia universalis. Lyon, 1684. Wiesemann, Claudia. “Instrumentalisierte Instrumente: EEG, zerebrale Angiographie und die Etablierung des Hirntodkonzeptes.” Instrument – Experiment. Historische Studien. Ed. Christoph Meinel. Berlin: GNT-Verlag, 2000. Willis, Thomas. Cerebri Anatome: cui accessit nervorum descriptio et usus. London, 1664. Willis, Thomas. “De Motu Musculari.” Opera Omnia. Vol. 1. Lyon, 1681. Willis, Thomas. “Two Discourses concerning the Soul of Brutes.” Dr. Willis’s Practice of Physick. Ed. Samuel Pordage. London, 1684. Willis, Thomas. “The Anatomy of the Brain.” The Remaining Medical Works of that Famous and Renowned Physician Dr. Thomas Willis. Trans. Samuel Pordage. 2 vols. Montreal, 1965 [facsimile of the London edition, 1681].
GERALD HARTUNG
The “Chymistry Laboratory”: On the Function of the Experiment in Seventeenth-Century Scientific Discourse This essay will deal with the “chymistry laboratory” as an exemplary arena for experimentation in the seventeenth century. However strangely some of the considerations presented might strike today’s reader, an “Entstehungsherd” (Friedrich Nietzsche) of experimental practice whose consequences reach into our own times will be assessed here. Since the introduction of the genealogical method into the writing of science history, we know that a practice and the frame of meaning of this practice do not form a stable unity. In our case this means that, although in the seventeenth and eighteenth centuries a radical shift in the method of interpretation took place within “chymical” science (at the threshold to a modern concept of the natural sciences, “alchemy” is vehemently criticized), we can still talk of a continual development and refining of experimental techniques. Thus the alchemistic experimental method turns out to be at the foundations of the modern chemistry that emerges in the eighteenth century. These various transitions are fluid, and the transition to the experimental culture of the modern period – this is the thesis of this essay – appears as a twisting path. For the theory and history of science, the central questions are: What is an experiment; how are “facts” produced in the course of an experiment; how is the relationship between these facts and their horizon of meaning constituted?1 Thus considering the performativity of the experiment and the career of experimental culture after the second half of the seventeenth century, it becomes possible to raise the question of the concept of scientific truth, without which the experimental method, to be considered here as a paradigm of modern science, could not have developed. 1
Cf. Steven Shapin and Simon Schaffer. Leviathan and the Air-Pump. Hobbes, Boyle, and the Experimental Life. Princeton: Princeton University Press, 1985. 3.
202
Gerald Hartung
1. Curiosity and Skepticism – Initial Remarks on the History of Science Why does the experimental method play a decisive role for the constitution of scientific truth? We are actually asking something self-evident, as almost nothing seems more obvious than the assumption that a process carried out in an experiment (repeatable at any moment) – e.g. the production of metal alloys – appears to be a fact with truth content. The true is what is actually verifiable for us as sensible and sensual beings. What is not available to the senses cannot claim to be the truth. A glance at the history of science reveals that this “attitude of mind” has not always been evident, and has its own history of emergence. Scholars have not always considered the experiment to be the ideal path to truth; and should this be the case, it does not necessarily follow that the truth content of a technical process – keeping with the example of metal alloys – is exhausted by what it is for us, or what is shown to us. For a long time, the sense of an impenetrable screen of deception limited man’s overestimation of his capacity for reason and most notions of self-empowerment, since “now we see through a glass, darkly . . . now I know in part” (1 Cor. 13.12). There was even a period of abstinence where any attempt at artificial imitation of nature, thus human curiosity, was considered reprehensible.2 The threshold between a culture that rejected experimental curiosity and the experimental culture of the modern period lies in the seventeenth century and is usually linked to the impact of a single scientific personality, “for historians are in wide agreement in identifying [Robert] Boyle as a founder of the experimental world in which scientists now live and operate.”3 There are several ways to counter such a scientifichistorical commonplace. It could be accepted, if it can be proved, that it is exaggerated, but not entirely false. There are however good reasons for not making it too easy. Apart from the fact that the history of science is not powered by single individuals, a contextual approach to the concept of modern experimental science is available. Steven Shapin and Simon Schaffer have incorporated the rational natural philosophy of Thomas Hobbes, and consequently confronted two competing concepts of rational science with each other, so that any preference for Boyle’s 2
3
Cf. Hans Blumenberg. “Curiosity is Enrolled in the Catalogue of Vices.” The Legitimacy of the Modern Age. Trans. R.M. Wallace. Cambridge: MIT Press, 1983. 309-25. Cf. Shapin and Schaffer. Leviathan. 5; Paul B. Wood. “Robert Boyle.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Jean-Pierre Schobinger. Vol. 3. Basel: Schwabe, 1988. 399.
The “Chymistry Laboratory”
203
concept over Hobbes’s cannot be grounded rationally, but must be understood politically. It is, however, also possible to confront the concept of modern experimental science with the – in a completely different sense – rational concept of nature which is characteristic of hermetic philosophy, which lies at the foundations of alchemy as well as the science of chemistry in the seventeenth century. I will take this path, and begin with a simple observation. Despite the above-mentioned rejection of experimental curiosity in Christian culture, there had still been a number of cultures of experiment – if not quite an experimental culture – before the seventeenth century. Experiments imitating, accelerating, or “improving” natural processes have existed since antiquity, for example with the flourishing of Babylonian, Egyptian, Greek, and Roman culture, to mention only these epochs within the Mediterranean area. There is plenty of evidence of practices of metallurgy, the production of “synthetic materials” – in the sense of counterfeiting valuable natural materials – and attempts at so-called “gold making.” Since the Middle Ages, these experimental practices have been grouped under one name derived from Arabic – “alchemy.”4 In the Christian Middle Ages too, alchemy held a great fascination. This is demonstrated by Albertus Magnus, Roger Bacon, Arnaldus de Villanova, and Thomas Aquinas.5 When, during the Renaissance, alchemy was transformed into a scientific discipline, this revaluation was accompanied by the creation of an ancestral gallery, whose motto “prisca sapientia” was meant to emphasize alchemy’s special dignity. Among its references are the wisdom of ancient Egypt, the Biblical creation myth (Genesis), the teaching of Hermes Trismegistos (Corpus hermeticum), the Arabic philosophy of nature (Tabula smaragdina), and cabalistic ideas, as well as Platonic, Aristotelian, Stoic, and Neoplatonic nat-
4 5
Cf. Hans-Werner Schütt. Auf der Suche nach dem Stein der Weisen. Die Geschichte der Alchemie. Munich: C.H. Beck, 2000. For extensive reference sources cf. Edmund Oskar von Lippmann. Entstehung und Ausbreitung der Alchemie. 3 vols. Berlin: Springer, 1919; John Read. Prelude to Chemistry. An Outline of Alchemy, its Literature and Relationships. London: G. Bell and Sons, Ltd., 1939; Elisabeth Ströker. Denkwege der Chemie. Elemente ihrer Wissenschaftstheorie. Freiburg and Munich: Alber, 1967. 71; Schütt. Stein der Weisen. For individual cases cf. A. George Molland. “Roger Bacon as Magician.” Traditio 30 (1974): 445-460; Stephan Meier-Oeser. “Roger Bacon oder Doctor mirabilis und die Macht der Wahrheit.” Zwischen Narretei und Weisheit. Biographische Skizzen und Konturen alter Gelehrsamkeit. Ed. Gerald Hartung and Wolf Peter Klein. Hildesheim: Olms, 1997. 95-124.
204
Gerald Hartung
ural philosophy.6 In the syncretism of Florentine Renaissance philosophy, these intellectual currents tend to merge, and by appeals to Christian authorities are protected against orthodoxy.7 In all these texts the talk is, directly or indirectly, of experimenting, of the success or failure of alchemical practices, above all of transmuting metals into gold. This current of thought is flanked by what was called “medical chemistry” introduced by Paracelsus in the early sixteenth century, and continued into the seventeenth century by his critics and followers Johann Baptista van Helmont and Johann Rudolf Glauber, who were practicing doctors, and who as alchemists were interested in the transmutation of matter.8 In this context the scholar Athanasius Kircher, who alongside his manifold projects also wrote alchemistic tractates, stands out.9 Though alchemy was generally the business of outsiders, it nevertheless stood at the center of scientific interests. It is not remarkable therefore that Robert Boyle showed no squeamishness about the alchemistic tradition, as is demonstrated, for instance, by his positive reception of the writings of van Helmont. Boyle’s famous treatise The sceptical chymist (1679) shows what effort at differentiation prepared the ground for the modern “culture of experiment.” Boyle acknowledged a considerable part of the experimental achievements of alchemy, rejected however its theoretical foundation. His skepticism was directed against the “philosophia hermetica” as the basis of the alchemistic experimental method, not against the latter itself. His intention – as his appeal to Francis Bacon and René Descartes and his critique of Cambridge Neoplatonism (in the person of Henry Moore) emphatically demonstrate – was to establish mechanical philosophy as the “accepted doctrine of the Royal Society.”10 This doctrine underlies our modern understanding of natural science. Here it is a matter only of those empirically verifiable natural facts whose structure and coherence is made visible in the experiment, as opposed to – as its 6
7 8
9 10
Cf. Wilhelm Schmidt-Biggemann. Philosophia perennis. Historische Umrisse abendländischer Spiritualität in Antike, Mittelalter und Früher Neuzeit. Frankfurt a.M.: Suhrkamp, 1998. Cf. Frances A. Yates. The Occult Philosophy in the Elisabethan Age. London and Boston: Routledge, 1979. 9-75. Cf. Wolf Peter Klein: “Franciscus Mercurius van Helmont.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Vol. 4.1. Basel: Schwabe, 2001. 48-53. Cf. Thomas Leinkauf. “Athanasius Kircher.” Holzhey and Schmidt-Biggemann. Philosophie des 17. Jahrhunderts. Vol. 4.1, 269-85. Cf. Paul B. Wood. “Die philosophische Ausrichtung der Royal Society.” Schobinger. Philosophie des 17. Jahrhunderts. Vol. 3.2, 391-94.
The “Chymistry Laboratory”
205
prehistory – the multiplicity of alchemistic strategies for supporting the thesis that the world is not only what is visible. In this context the experiment does not aim at unlocking nature’s secret and making this visible. Neither in the Paracelsian mode of natural philosophy, nor in the various currents of thought linked to the concept of a “prisca sapientia” is it a matter of active visualization, but rather of the notion that the observation of nature is separated from the core of knowledge by an unsurmountable gap, and that the latter can only be conveyed to us through the medium of revelation. What is visible, “the surface – of texts as well as things – is understood as the manifestation which veils an actual, inner, spiritual content.”11 This is not at all contradicted by the fact that in the seventeenth century the ideal of the “experientia” begins to prevail. This can be thought without contradiction so long as the hierarchy of knowledge remains intact, i.e. empirical science is understood as the applied discipline whose highest perfection lies in the practice of a “magia naturalis” or “alchymia.” It could be, however, as formulated at the beginning, that the definition of practice and the method of interpretation have shifted. Consequently, the situation could arise that an experimenter as practitioner is in advance of his times, although – or precisely because? – he sticks to a traditional method of interpretation. This is the case with one of the greatest experimenters of the late seventeenth century who, depending on one’s point of view, could be seen as the last great alchemist of his time or as one of the founders of scientific chemistry. 2. Project Making and Experimental Method – The Alchemist Johann Joachim Becher Anyone who deals with the experiment, the theoretical foundation of experimentation, and the question “Why experiment at all?” in the early modern period must cast at least a cursory glance over the figures who are practicing. Here one often comes across the type of early modern polymath incorporating philosopher, theologian, and naturalist. This figure generally does not belong to the academic tradition, does not teach at the university; nevertheless, he decisively determines the dynamic of the scientific discourses of his times. This was the case with Descartes, 11
Stephan Meier-Oeser. “Einleitung zu § 1. Hermetisch-platonische Naturphilosophie.” Holzhey and Schmidt-Biggemann. Philosophie des 17. Jahrhunderts. Vol. 4.1, 11.
206
Gerald Hartung
Hobbes, Baruch de Spinoza, John Locke, and Gottfried Wilhelm Leibniz – and this was also the case with most natural philosophers from Roger Bacon to Robert Boyle and his circle, who founded their own institution – the Royal Society. What these scholars have in common is partly the manner of their intellectual development, which was the result of self-acquired knowledge rather than knowledge received from the appropriate institution. It should be added that the lack of connection to institutions as well as the social and political instability of their circumstances was so marked that their research and work had the character of project making and were limited by the restraint of practicability. The latter thesis is well demonstrated, in my opinion, by the life and work of the polymath Johann Joachim Becher, who gathers together the main tendencies of his times. He stylized himself as “self-taught” and a “project maker”; the next generation saw him as an example of the eclectic scholar who examined every theoretical insight for its practical use, and therefore as a model of the “useful scholar.” Later, however, his reputation was distorted into the image of the alchemistic charlatan.12 The fact is that throughout Europe he worked in various alchemistic laboratories, including the Court of Munich, where in this period, “the finest chemical laboratory in Europe”13 could be found. His work included tractates on philology, moral philosophy, economics, psychology, medicine, and alchemy.14 Becher himself at one point mentioned the difficulties connected with the title of alchemist in the seventeenth century: Since out of natural inclination, and therefore as a result of divine vocation, I am required to inquire into natural things, and so from the time I could barely walk I strove through various texts on chemistry to become familiar with the world, I could not help, for the reasons mentioned, being called an alchemist, a name that clung to me until I could manage to shake it off by really dedicating myself to alchemy’s concerns as well.15 12
13 14
15
Cf. Hermann Kopp. “Die Alchemie bis zum letzten Viertel des 18. Jahrhunderts.” Die Alchemie in älterer und neuerer Zeit. Ein Beitrag zur Kulturgeschichte. Hildesheim: Olms, 1962 [facsimile of the edition Heidelberg, 1886]. 66ff.; James Riddick Partington. A History of Chemistry. Vol. 2. London: Macmillan, 1961. 638; Gerald Hartung. “Johann Joachim Becher.” Holzhey and Schmidt-Biggemann. Philosophie des 17. Jahrhunderts. Vol. 4.1, 190-99. Lynn Thorndike. A History of Magic and Experimental Science. 8 vols. New York: Columbia University Press, 1964. Vol. 7, 579. Cf. Pamela H. Smith. “Chemistry and Commerce. Johann Joachim Becher and the Court of the Elector in Munich.” Johann Joachim Becher (1635-1682). Ed. Gotthardt Frühsorge and Gerhard F. Strasser. Wiesbaden: Harrassowitz, 1993. 143-58. “Dann weil ich auß natuerlicher Zuneigung/ und also durch eine Goettliche Voca-
The “Chymistry Laboratory”
207
On alchemy in the narrow sense he published a whole series of writings, which can be roughly separated into two categories: 1. theoretical writings, natural philosophical treatises and 2. practical writings, presentations of his experimental method. Belonging to the first category is his first alchemical work, the Natur-Kündigung der Metallen (1661),16 with the Supplementa,17 as well as the Oedipus chimicus (1680).18 Among the practical writings belong the famous Experimentum chymicum novum (1671),19 which treats the generatio of metals, the Trifolium Hollandicum (1679),20 the Chymische Laboratorium (1680)21 and the treatises Chymische Glückshafen (1682)22 and Chymische Rosen Garten (1717),23 which contain a compilation of experiments and a description of the important laboratories in Europe. On the threshold between theoretical speculation and a practical guide there is a noteworthy text, the Tripus hermeticus fatidicus (1689),24 which combines theo-
16 17 18 19
20
21 22
23
24
tion, den natuerlichen Sachen nachzuforschen gehalten bin/ und gleichsam von Kindes Beinen an/ mich darauff beflissen/ mich auch der Welt durch mancherley Chymische Schriften schon allbereit bekandt gemacht/ habe ich mich nit hueten koennen/ daß mir/ auß besagte Ursachen/ nicht auch die Klette eines Alchymisten angehenkt worden/ welche ich zwar anders nicht abschuetteln unterstehen koennen/ als wan ich mich der Alchymie ihrer Sach mitannehmen wuerde.” Johann Joachim Becher. Experimentum Chymicum Novum: Oder Neue Chymische Prob, Worinnen Die künstliche gleich=darstellige Transmutation, oder Verwandelung/ derer Metallen/ augenscheinlich dargethan. Frankfurt, 1680. 11. Johann Joachim Becher. Natur=Kuendigung der Metallen. Mit vielen Curiosen/ Beweißthumben/Natuerlichen Gruenden/Gleichnuessen/ Erfahrenheiten. Frankfurt, 1661. Johann Joachim Becher. Supplementum secundum in physicam subterraneam, id est . . .. Frankfurt, 1675. Johann Joachim Becher. Institutiones chimicae prodromae, id est J.J. Becheri Spirensis Oedipus chimicus. 2nd ed., Frankfurt, 1680. Johann Joachim Becher. Experimentum chymicum novum, quo artificialis & instantanea Metallorum Generatio & Transmutatio ad Oculum demonstratur. Frankfurt, 1671. Johann Joachim Becher. Trifolium Becherianum Hollandicum oder . . . Dr. Joh. Bechers Drey Neue Erfindungen Bestehende in einer Seiden=Wasser=Muehle und Schmelz=Wercke. Frankfurt, 1679. Johann Joachim Becher. Chymisches Laboratorium Oder unter-erdische Naturkuendigung. Frankfurt, 1680. Johann Joachim Becher. Chymischer Glücks-Hafen oder große Chymische Concordanz und Collection/ Von fuenfzehen hundert Chymischen Processen. Frankfurt, 1682. Johann Joachim Becher. Chymischer Rosen=Garten, Samt einer Vorrede und kurtz gefassten Lebens=Beschreibung Herrn D. Bechers. Ed. Friedrich RothScholtzen. Nuremberg, 1717. Johann Joachim Becher. Tripus hermeticus fatidicus, pandens oracular chymica. Frankfurt, 1689.
208
Gerald Hartung
retical considerations with practical suggestions, as I shall go on to show, in a particularly striking manner. A glance at the title of the treatises as a whole shows the two subjects that are dealt with prominently. Firstly, the concept of hermetic philosophy, whose tradition is repeatedly referred to, is retained; secondly, however, the term “experimentum” becomes appropriate for a title, and the linking of natural causes, similarities, and “experiences” in the course of giving evidence, and for the transformation of the metals, “ad oculum demonstrare,” are promised. This raises the question of how experimentation functions within Becher’s natural-philosophical conception. How close does the observable demonstration of nature’s acts come to its secret? What is important here: veiling or unveiling? 3. Experiment and Secret – On the Function of the Experiment in Alchemy The internal correspondence of “experiment and secret” is founded on the conception of hermetic philosophy. One can observe this correspondence most clearly in the work of Kircher, whose writings constantly represent a point of orientation for Becher’s own studies to the extent that the titles of a few of his writings – the Oedipus chimicus above all – brought accusations of plagiarsm. Kircher had developed his idea of a universal science from principles of ancient wisdom, Neoplatonic philosophy, as well as Lullian methodology. Here, the universe is considered from the point of view of unity and organic interaction; everything is meaningfully ordered, and from a single – conceivably possible – divine perspective everything is perfect as it is.25 The fact that we perceive things differently is not only our all-too-human problem, but it is also of its own making. Only after the Fall was an appeal to empirical data necessary since the ideal of purely contemplative, immediate knowledge of objects had been lost. This is the theological basis of the early modern conception of science, according to which turning to the knowledge of facts was due to “the insight into the theological cause of the disempowerment of human powers of perception.”26 In his tractate Mundus subterraneus, Kircher states that the experientia can lead us from the analysis of the next cause (causae proximae) to the knowledge of the “interior of things.” This position suggests two inferences: 1. Empi25 26
Cf. Leinkauf. “Athanasius Kircher.” Ibid. 274.
The “Chymistry Laboratory”
209
rical knowledge is now the ideal path to the perception of truth. 2. Empirical knowledge – and for Kircher this also included the observation of nature in an experiment – is useless in the search for truth if it is not supported by the principles of natural philosophy. Here, the connection between the two positions is more than questionable, since the validity of the principles of natural philosophy – cosmological unity and organic interaction – does not depend on the fact that they are confirmed in the observation of nature or in the experimental procedure; instead any divergence is then explained theologically. Whether the materiality of things as well as the structure of organic processes is made visible to us depends on the inferences we draw from observation; and for this, in the last instance, theology is responsible. The embrace of experimental practices remained half-hearted. “Kircher makes a great show of experimental method, asserting that to philosophize in physical matters without experiments is the same as if a blind man should presume to be judge of colors . . . But his own experiments seem silly and insufficient.”27 We will see that Becher, though largely standing within this tradition, goes beyond it in a few decisive places.28 In the year 1661 he published his first alchemistic treatise with the title Zur Natur=Kuendigung der Metallen. This text discusses the basic problem of alchemy, i.e. the imperfection of the majority of metals compared with those considered precious, particularly gold, and the proof that alchemy has not only the theoretical conditions, but also the practical ability to overcome this lack. Thus the alchemist is to the metallic what the doctor is to the animal body; both submit what is imperfect to a therapy, i.e. the alchemist attempts to heal “the internal malady of the imperfect metals.”29 Becher is fundamentally committed to the principles of unity and the structural coherence of the natural order. He divides this – entirely traditionally – into three natural realms and one artificial realm. The first, the animal, vegetable, and metallic realms, belong to nature and are produced by it. In this division into three parts, the imperfection of nature is already visible, since the Trinity represents, according to the criterion of a creation from one divine origin, a defect, which is therefore seen to be in need of redemption. What marks the Fall in terms of the theology of creation and salvation can be seen from the natural philo27 28
29
Thorndike. History of Magic. 571. The Actorum Laboratorii Chymici Monacensis seu Physicae Subterraneae libri duo published 1669, four years after Kircher’s text. On the intellectual historical background cf. ibid. 578-90. Becher. Natur=Kuendigung. 17.
210
Gerald Hartung
sophical perspective as a development. Both aspects relate to each other as a pattern for the explanation and description of one and the same thing. Only under particular (theological) conditions does the description of nature make sense at all. Generatio, the basic concept of Aristotelian natural philosophy, is transferred by Becher, by means of analogy, from the organic world (animal-vegetable) to the inorganic world of metals. It is thus shown what targets alchemy aims at in theory and practice. Alchemy’s task is to understand the generation of metals, but this task exists in a larger context. The understanding of generation is the condition for its surpassing [Aufhebung], since this is the aim of alchemy: to promote the redemption of the metallic realm just as all “natural sciences” take part in the redemption of the whole natural order. Proof that Becher’s approach is also shared by other authors and should not be rashly logged under the curiosa of the history of thought is given by Isaac Newton’s treatise Vegetation of Metals (1669) where he speculates on the universal force of nature, its secret fire.30 Along with these three realms, of which the alchemist mainly deals with one, there is, according to Becher, a fourth realm. Unlike the other realms, this is not a natural but an artificial product – it arises from human intervention. The fourth realm first arises in the experiment, for example as a result of the separation and mixing of metallic substances, and as a result of a process of combustion giving rise to the transmuting of ores.31 Here the shift in meaning is already revealed. Human intervention in the natural order, the experiment, not only confirms what essentially takes place, but is itself the medium of this event. A glance at the structure of the Natur=Kuendigung der Metallen makes this connection clear. Here Becher progresses from part one (“Conception and Birth of the Minerals”) dealing with the origin of minerals and metals and the process of generation, through part two (“Smelting and Release of the Ores”) explaining how, by means of separation, the alchemists might prevent the mixing and contamination of metals, to part three (“Transmuting of Metals”) where the constitution of the fourth realm is completed. In other words it describes the emergence of the metallic realm as a process of mixing, going on to explain the conditions of unmixing or separation, to finally deal with the transmutation or perfection of the metals. 30 31
Cf. Lorraine Daston. Eine kurze Geschichte der wissenschaftlichen Aufmerksamkeit. Munich: C.F. von Siemens Stiftung, 2000. 15. Cf. Becher. Experimentum Chymicum.
The “Chymistry Laboratory”
211
How can this be imagined? Becher’s alchemistic assumption is that all subterranean bodies (metals) are of an earthy nature. Here, a distinction should be made between the vitrifiable, combustible, and mercurial basic earths. These contain the alchemistic principles of dissolvability, combustibility, and volatility. All metals contain these three elements; only the proportion differs. The transmuting of metals is possible by means of the combustion of the combustible earths. During combustion, a volatile, mercurial, and dissolvable fluid material appears, which stands in some – still to be determined – relation to first matter. Here, Becher takes up an old theme of alchemy: the creation of “prima materia” and the search for the so-called “philosopher’s stone.” First matter is – already in Aristotle’s definition – the basic material without qualities, which manifests itself in the various forms of the natural body, though usually hidden from sight. If for example a metal is returned (in the combustion process) to this original state, it becomes receptive again (as at the beginning of creation) to all kinds of forming, or to the adoption on of qualities. The “philosopher’s stone” can now be understood as such a power, which forms material in an ideal sense; in the case of metals, re-forming them as precious metals. As a prerequisite so that metals transmute into gold there must be a material basis having the consistence of mercury, a quality however opposed to normal mercury. This basis is called tincture or the philosopher’s stone. A tincture: because it colors the metals gold. Generally however a more elegant designation is derived from the color as a superficial accidental thing. It is called Lapis Philosophorum because the philosophical chemists intend, with the tincture, to produce such a thing.32
The general external description of this tincture comes very close to that of mercury; its qualities, however, should be incomparable. It is not a question of hard, fluid, or volatile material, thus ruling out mercury and ether, but of the principle of generation per se. In fact this is a matter of the principle of the optimized and temporally compressed artificial generation of metals as opposed to natural generation. “Becher was totally convinced that the transmutation of metals in general and the 32
“Das Obhaben/ so die Metallen in Gold verwandelt/muß seyn ein materialisch Unterhaben/ die eines Bestandwesens des Quecksilbers/ der Eigenschafft aber/ dem gemeinen Quecksilber entgegen. Diß Unterhaben wird die Tinctur/ oder der Stein der Weisen genennet. Eine Tinctur; weiln es die Metallen in Gold färbet. Es wird aber/ insgemein/ die vornehmere Benambsung genommen von der Farb/ als einem/ in das Gesicht fallenden/ zufälligem Dinge. Lapis Philosophorum wird es genennet/ weiln die philosophische Chymisten/ mit der Tinctur/ dieselbe zu verfertigen/ zu schaffen haben.” Becher. “Nochmaliger Zusatz über die Unter-erdische Naturkuendigung.” Becher. Chymisches Laboratorium. 33.
212
Gerald Hartung
transmuting of metals into gold in particular was possible, and that the universal method of bringing about the latter could actually be demonstrated.”33 To make his idea plausible, the alchemist argues thus: God, as the origin of everything that exists, is also the source of the transformation of everything that materially exists; and this is also the principle of the natural emergence of gold by the generation of metals, and the condition that permits this process to be emulated in the alchemistic experiment. Apart from the fact that the conditions of the experiment remain secret – the production of the “lapis philosophorum” results from the process of the “cooking of something with mercury” – the experiment is intended to prove to the senses that the transformations in the processes of nature, and the principle of perfecting, are contained within it. In the case of gold-making – as artificial imitation of the coming about of gold, the special case of the generatio – it is a matter of the proof that all metals have a tendency to be transmuted into gold, “thus he proved that all metals strive for the state of gold.”34 Seen in this way, the tincture merely has the task of affecting the respective non-precious metal with an aspect of perfection, to elicit so-called innate gold [Goldwesen]. The first to know how to work on the tincture is Mercurius who prepares the gold and makes it subtle and penetrable, so that it can tint. Then the tincture makes gold because it is gold; it expands however its gold essence into an innumerable duplication, because it is made subtle and penetrated. And however in no way can such a making subtle happen without a reconcentration by means of Mercurius.35
Subtracting all the inconsistencies and secrecy, we can reasonably say that Becher’s search for the first matter of Aristotelian physics is orientated in its intentions towards practically executable experiment. It turns against the “fantasies of the cabalists,” who actually want to make material emerge and imitate the godly creation ex nihilo. This is, as he often stresses, the prerogative of God alone; the human attempt to emulate “creatio ex nihilo” would be an inexcusable blasphemy. In this sense Becher brings his critique of the Aristotelians and the magi33 34 35
Kopp. Alchemie. 67. Becher. Experimentum Chymicum. 40. “Der Anfang/ die Weiß nemlich/ an der Tinctur zu arbeiten/ ist der Mercurius, der das Gold aufschliest/ subtilirt/ und durchdringend macht/ deß es tingiren koenne. Dann die Tinctur macht Gold/ weil es Gold ist; erstrecket aber sein Goldwesen in einer unzehlbaren Vervielfaeltigung/ weil es subtilirt und durchdrungen ist. Und kan aber eine solche Subtilirung/ ohne die Wiederzusammenpunctung/ keines Wegs/ als durch den Mercurium, geschehen.” Ibid. 42.
The “Chymistry Laboratory”
213
cians/cabalists together. The former wrongly consider a “reduction of the metals to their matter” impossible because under the first matter of the latter they understand an “ens rationis” that is nothing other than the highest folly. The “chymici” however, among whom Becher includes himself, understand something different by the reduction of the metals to a first matter. Anyone can see “what the chymici understand by a reduction to first matter, which they call first matter because it is the first before our eyes.”36 He also speaks of his project as a “chymical philosophy” which does not aim at the transformation of something into nothing, but where material bodies are simply combusted.37 In addition it should be said that in the experiment the mixing ratios of the hard, fluid, and volatile earths are reconstructed as a formal principle by the addition of the “prima materia,” thus the actual state of creation is optimized. Alchemy or “chymical philosophy” is therefore seen foremost as an experimental method based on the processes of separation and mixing in the combustion process. In this process the elements mixed in a natural way are separated (analysis) eventually to be combined together artificially in a particular mixing ratio (synthesis). Combination in the experiment closes the gap between observed (empirical) knowledge and the non-visible ordering principles of nature.38 The success of the alchemistic method essentially depends on the attitude of the experimenter, since the reason that something succeeds or an experiment is successful – e.g. the creation of gold by the smelting of sand39 – depends essentially on the ability to bring standards and philosophical principles, the visible and invisible world into union, without, by overestimating one’s own abilities (through magical practices), denying the existing differences. It is not a matter of unveiling the divine secret of the natural order, but – paradoxically – of humility before the unfathomable secrets of the acts of nature in the experiment of intervening in an optimizing way in the processes of nature to optimize the generation of metals.40 36
37 38 39 40
“Was die Chymici durch das Reduciren/ in ihre erste Materi verstehen/ welches sie darumb die erste Materi nennen/ dieweil sie die erste in unsern Augen ist.” Becher. “Doctor Bechers Philosophisches Gutachten ueber die Philosophische Tinctur im trucknen Wege. Fünf Artikel.” Becher. Chymischer Glücks-Hafen. 36. Becher. Institutiones chimicae. 11. Thorndike. History of Magic. 137. Becher. Chymisches Laboratorium. 731. Cf. Henricus A.M. Snelders. “Johann Joachim Becher und sein Gold-aus-Sand-Projekt.” Frühsorge and Strasser. Johann Joachim Becher. 103-14. Biographical information on Johann Joachim Becher and the history of his works
214
Gerald Hartung
4. The Laboratory – Structure and Conception of the Experiment From prominent quarters Becher received the accusation that he was an intelligent researcher, but that precisely that “very little use can be made” of his most important book, the Natur=Kuendigung der Metallen. The most important chemist of the early eighteenth century, Georg Ernst Stahl, the founder of the phlogiston-theory,41 emphasized that Becher tempts the reader of his texts “as though blindly into the making of gold and silver.”42 Nevertheless, he had published a paper of Becher’s himself, and called him a precursor of scientific chemistry.43 Becher was also aware of the danger of blind emulation resulting from the ambivalence of an alchemistic science of principles and an empirical knowledge oriented towards experiment. For this reason he pleaded emphatically for a thorough study: To learn science from its foundations it is necessary to look at its principles and the resulting axioms that come from observation and experiment; and these have their origin in the combinations and concordances, and this is the true and proper methodus in all sciences.44
To learn the “methodus” is the main aim of a study which a scholar can only achieve in cooperation with other colleagues. As a result, it is also clear, that many connoisseurs and artists, laboratories, works, and experiments were needed and many great men to whom I often offered my services on my travels so that I could obtain the journal of experiments that they had carried out in their laboratories.45
41
42 43
44
45
is given in Gerald Hartung. “Johann Joachim Becher oder: Die Projekte und Konzepte eines närrischen Gelehrten.” Ibid. 262-87. Cf. Partington. History of Chemistry. 653 and Elisabeth Ströker. Theoriewandel in der Wissenschaftsgeschichte. Chemie im 18. Jahrhundert. Frankfurt a.M.: Vittorio Klostermann, 1982. 78. Georg Ernst Stahl. Bedencken, Erinnerung und Erlaeuterung über Doctor J. Bechers Natur-Kuendigung der Metallen. Frankfurt and Leipzig, 1723. 442. “The future influence of Becher’s Subterraneam Physics is attested by further editions of it at Leipzig in 1703 and 1738. In both cases an addition by G.E. Stahl speaks of Becher’s book as ‘a work without an equal.’” Thorndike. History of Magic. 583-84. Cf. Ströker. Theoriewandel. 83-84. “Nun die Wissenschafften recht auß dem Fundament zu lernen/ ist noethig auf ihre Principia und hierauß erfolgende Axiomata zusehen/ welche auß den Observationen und Experimenten kommen: und diese haben wiederum ihren Ursprung auß den Combinationen und Concordantien/ und dieses ist der warhafftige/ meisterliche Methodus in allen Scientien.” Becher. Chymischer Glücks-Hafen. “Vorrede,” I. “Daß vielerhand Liebhaber und Künstler/ Laboratoria, Arbeiten und Experimen-
The “Chymistry Laboratory”
215
Becher produces as evidence in his Chymischen Glückshafen (1682) a list of the manuscripts and laboratories from which he claims to have developed his concordance: the laboratories of Kaiser Rudolf II, Ferdinand III, of Friedrich, the Count Palatine of Mainz, the Kur-Kölln as well as the Holstein Laboratory, etc. Anticipating Stahl’s objection that the presentation of such a quantity of chymical processes presents a danger, I answer him that this book or chemical concordance was not written for just any idiot, but for experienced laboratory workers, who can decide themselves from the concordance what they need to do and what they want and should undertake to work on.46
The character of alchemistic experimentation, which depends essentially on the attitude of the experimenter and is aimed neither at the public nor at transparency, brings along an abundance of difficulties stemming from the aforementioned ambivalence. This expresses itself in the form of the experiment in the opposition of esoteric and exoteric approaches, in Becher’s words, a “methodus gnostica” and a “methodus didactica.” The latter has two main aspects, to which Becher attaches great importance: the “methodus laborandi” and the instructions for the construction of a laboratory. Becher’s presentation of the “methodus laborandi” is based on an extensive enumeration of all possible errors that can appear during experiments, generally to be attributed to human weakness. So there are a few who have everything done by the hands of others and who therefore must believe what their laboratory workers tell them, and those are usually quite mean ignorant scum, liars, and drunkards, generally shady rascal, who also are easily bribed by others, and for a taste of wine they trade their work, that has cost their masters so much money, people, and trouble.47
In fact it is a matter here, after we have left the Gnostic aspect of the acquisition of the right approach behind us, of the question of the prac-
46
47
ten darzu erfordert wuerden/ derentwegen ich mich auf meinen Reisen und vielerhand grosser Herren Diensten allzeit bemuehet/ daß ich die Journal von Experimenten/ die sie in ihren Laboratoriis gethan/ haben koennen.” Ibid. III. “Dem gebe ich zur Antwort: daß dieses Buch oder Chymische Concordanz nicht vor einen jeden Idioten/ sondern vor erfahrene Laboranten geschrieben sey/ welche auß der Concordantz selbst judiciren koennen, was sie zu thun haben und darauß zu laboriren vornehmen wollen und sollen.” Ibid. V. “So seyend auch einige/ welche alles durch frembde Haend lassen arbeiten/ derentwegen glauben muessen/ was ihnen ihre Laboranten vorsagen/ welches gemeiniglich unverständige Kohlenschürer/ verlogene und versoffene/ ja ins gemein verstohlene Tropffen seynd/ welche auch umb ein leichtes von andern sich bestechen lassen/ und ihre Arbeit umb einen Trunck Wein communiciren/ die doch ihren Herrn so viel Geld/ Leuth/ und Muehe gekostet hat.” Ibid. 74.
216
Gerald Hartung
tical conversion of a theoretical concept into a particular sequence of actions. The aim is to convey the truth, to maintain the order, and safeguard the secret of the successful experiment. “I have therefore undertaken to show in this article a method of practicing alchemy usefully and with pleasure so that there is neither delusion nor deception and everything can be kept secret and in good order.” And he lists five useful rules, the application of which will bring about success. The first three are as follows: Then 1. all works are carried out philosophically because they are firstly considered and pondered, since nothing can be known of such difficult processes. 2. Everything will be carried out in an orderly fashion: neither the process nor the operations, nor the materials, nor the instruments, nor the people will come in contact, mix together, confuse each other, collude, confer, communicate, and, as it is customary amongst laboratory workers, drink alcohol. 3. Everything will follow a perfect order since the consiliarii will constantly be occupied with their studies, the dispensator with the distribution of his operations, the other 3. laboratoria with their own work, and since nobody has several tasks, but each always only one, and stays with one operation, then it cannot fail that all parties become perfect in their operation.48
Becher reveals himself as an experienced experimenter discussing the advantages and disadvantages of a division-of-labor approach, and considers a whole series of control measures to limit the defectiveness of human nature as the main risk factor in the field of research. The construction of the laboratory is described in detail in Becher’s Tripus Hermeticus Fatidicus. Its first essay delivers a plan for the construction and assembly of a Laboratorium Portatile. The schematic presentation of the elements – measuring instruments, various types of com48
“Ich hab mir derohalben vorgenommen/ in diesem Articul einen Methodum zu weisen/ wie man die Alchymi nutzbarlich/ und mit einem Lust/ practicè tractiren/ darinnen nicht betrogen/ noch belogen/ auch alles geheimb/ und in guter Ordnung gehalten werden koenne . . . Dann I. werden alle Arbeiten/ so vorgenommen werden/ Philosophisch zu gehen/ dieweil sie erstlich ueberleget/ und ponderirt werden/ ist sich derohalben von so ungeraeubmten Processen nicht zu befahren. 2. Wird alles ordentlich zugehen: Es werden weder die Process/ noch die Operationes/ noch die Materialien/ noch die Instrumenten/ noch die Menschen zusammen kommen/ sich vermischen/ confundiren/ miteinander colludiren/ conferiren/ communiciren/ und was bey den Laboranten gebraeuchlich/ sauffen koennen. 3. Wird alles in einer Perfection hergehen/ dann die Consiliarii werden staettigs in ihren Studiis begriffen seyn/ der Dispensator mit Außtheilung seiner Operationen/ die uebrige 3. Laboratoria mit ihren eigenen Arbeiten/ und dieweil keiner viellerley/ sondern allzeit jeder nur einerley thut/ und bey einerley Handlung verbleibt/ so kan nicht fehlen/ es muessen auch alle Classen in ihrer Operation perfectioniren.” Ibid. 102.
The “Chymistry Laboratory”
217
bustion furnaces and various test materials – is completed by the Alphabetum minerale, which in Lullian form contains philosophical guidance and a classification schema of the mineral elements.49 Becher’s plan for a portable laboratory is unique in its time and equally a landmark. Becher’s book on the portable furnace has been considered because of the widely recognized problem of inflexibility of the traditional masonry type of furnace, some being incorporated into the fabric of the laboratory . . . Becher designed a furnace that was both portable and versatile. Each chemical process – fusion, cupellation, calcinations, reverberation, cementation, distillation, digestion (and the bain-marie), sublimation, fierce distillation, and distillation per ascendum – needed a particular kind of furnace. Becher’s simple solution was to provide an eight-piece kit of parts that could be put together in six ways. The furnace was made of earthenware.50
Above all, its versatility – enabling all chemical processes from smelting to distillation and combustion – made this invention an excellent achievement. The laboratory was actually built following Becher’s plan and went into production, though not in his lifetime, but in the form of a copy by Peter Shaw who in 1731 published his Essay for Introducing a Portable Laboratory, going on to sell the idea in collaboration with the instrument maker Francis Hauksbee.51 This latter episode also sheds light on the question of the effect of Becher’s alchemistic experimental method. Here we see an example of the selective use “scientific” chemistry makes of its precursor, alchemy or “chymical philosophy.” It adopts the classification of elements and the instructions for the experimental process, and also the process of combustion (without its theoretical foundation), and developed from this is the modern combustion theory – i.e. the Stahlian phlogiston theory.52 Stahl’s definition of chemistry does not, therefore, outwardly differ from Becher’s definition of chymical philosophy: “Chemistry, which is also called alchemia and spagyrica, is an art which dismantles mixed and put or piled together . . . bodies into their principia, or which reunites from such principii to the same body.”53 49 50
51 52 53
Becher. Tripus hermeticus. 1-149. Robert G.W. Anderson. “The Archaeology of Chemistry.” Instruments and Experimentation in the History of Chemistry. Ed. Frederic L. Holmes and Trevor H. Levere. Cambridge: MIT Press, 2000. 17. Cf. ibid.17-18. Cf. Ströker. Theoriewandel. 87. “Chymia alias Alchymia et Spagirica, et ars corporea vel mixta vel composita, vel aggregata etiam in principia sua solvendi aut ex principiis in combinandi.” Georg Ernst Stahl. Fundamenta Chymica Dogmatico Rationalis et Experimentalis. Nuremberg, 1732. 6. Quoted in Ströker. Theoriewandel. 89.
218
Gerald Hartung
The differences between alchemy or “philosophical chemistry” on the one hand, and chemistry as a scientific discipline on the other are not founded on the practice of experimentation itself, but result from a shift in the method of interpretation. This shift can be understood from the following considerations: 1. In the experiment, the principles of the metallic body are revealed. These are the elements that cannot be traced back to original substances (water, earth, etc.), but are subsumed according to their functional performance (fluid, combustible, etc.). Becher attempted to unite both – the substantial and the functional definition. Stahl, however, limits the experimenter’s gaze to revealing purely functional connections. 2. The experiment as artificial technical process of separation and mixing is completely removed from the speculative horizon by Stahl. From Becher’s double perspective, which considers the “methodus gnostica” as well as the “methodus didactica,” only the exoteric path remains; chemistry becomes a merely empirical and applied science. 3. The experiment makes visible what is visible for us as observer. The fact that not everything becomes visible is for Stahl merely due to pragmatic considerations (refinement of the technical apparatus) and no longer has, as it still does with Becher, the character of a secret. 4. The experiment aims at verifiability, and in this way abandons the character of a secret. The experimenter appears to be fundamentally interchangeable, to the extent that the difference between spiritual (methodus gnostica) and practical (methodus didactica) experience is leveled out. These points, and in fact each one separately, demonstrate the possibility that a famous experimenter might be ahead of his time as a practitioner, although – and precisely because! – he held firm to an inherited method of interpretation. What consequences does this thesis bring to the genealogical definition of the origins of experimental culture? 5. From “Alchymia” to “Chymie” – Remarks on the Genealogy of Experimental Culture in the Seventeenth Century The restricted view of the role played by the experiment in the discourse of alchemy, well represented by the writings of the polymath Becher, strengthens – with the reservation that it is of limited expressiveness – the initially formulated thesis. This snapshot of a moment in the history of science underscores our supposition that the path to the experimental culture of the modern period, in terms of its beginnings in
The “Chymistry Laboratory”
219
the seventeenth century, cannot assume a straightforward development. Beyond the debate between Boyle and Hobbes on “the polity of science,” between naturalists in the alchemistic tradition and the new generation of empirical scientists, an almost noiseless shift in meaning takes place. In this process, Becher, the “chymical philosopher,” assumes a notable position because he 1. reveals the ambivalence of this process in its double – namely esoteric and exoteric – perspective, and 2. thus demands a shift in meaning that enables a large measure of continuity in the practical procedures of “chymical” experimenting. If the description of an experiment carried out under laboratory conditions sufficiently reveals the natural event to be depicted, then it would need no further explanation; for those which merely depend on its sensible experience, any explanation appears to be delivered a posteriori and not, in principle, to be given. Therefore Becher describes the technical implementation and the functional description of the individual parts of the process of combustion in a way that his successor can simply adopt while at the same time renouncing the pattern of explanation. This consideration provokes the thesis that the shift in meaning in relation to natural objects has a long-term scientific-historical, and that means also ontological, resonance. While experimental practice changed only gradually and was continuously refined, it has been shown that its method of interpretation was based on a radical change. What function the experiment has, and what is shown in the experiment, is linked to historical change. If the logic of being is subjected to the logic of changing perspectives, then ontology has a history. Or in the words of Lorraine Daston, it might be that from a divine point of view the inventory of the universe, the fullness of all that exists is always and everywhere the same, but if one observes it from the perspective of what objects of research are approached by the sciences, then ontology has a history.54 To which it can be added that not only the changing approaches to the objects, but the ever-changing manner of the approach present a great challenge to the writing of the history of science. With the history of ontology the historicity of scientific perspectivism emerges. Translation: Benjamin Carter
54
Cf. Daston. Aufmerksamkeit. 11.
220
Gerald Hartung
WORKS CITED Anderson, Robert G.W. “The Archaeology of Chemistry.” Instruments and Experimentation in the History of Chemistry. Ed. Frederic L. Holmes and Trevor H. Levere. Cambridge: MIT Press, 2000. Becher, Johann Joachim. Natur=Kuendigung der Metallen. Mit vielen Curiosen/ Beweißthumben/ Natuerlichen Gruenden/ Gleichnuessen/ Erfahrenheiten. Frankfurt, 1661. Becher, Johann Joachim. Experimentum chymicum novum, quo artificialis & instantanea Metallorum Generatio & Transmutatio ad Oculum demonstratur. Frankfurt, 1671. Becher, Johann Joachim. Supplementum secundum in physicam subterraneam, id est . . . Demonstratio Philosophica, seu Theses chymicae, veritatem & possibilitatem transmutationis metallorum in aurum evincentes. Frankfurt, 1675. Becher, Johann Joachim. Trifolium Becherianum Hollandicum oder . . . Dr. Joh. Bechers Drey Neue Erfindungen Bestehende in einer Seiden=Wasser=Muehle und Schmelz=Wercke. Frankfurt, 1679. Becher, Johann Joachim. Chymisches Laboratorium Oder unter-erdische Naturkuendigung. Frankfurt, 1680. Becher, Johann Joachim. Experimentum Chymicum Novum: Oder Neue Chymische Prob, Worinnen Die künstliche gleich=darstellige Transmutation, oder Verwandelung/ derer Metallen/ augenscheinlich dargethan. Frankfurt, 1680. Becher, Johann Joachim. Institutiones chimicae prodromae, id est J.J. Becheri Spirensis Oedipus chimicus. 2nd ed., Frankfurt, 1680. Becher, Johann Joachim. Chymischer Glücks-Hafen oder große Chymische Concordanz und Collection/ Von fuenfzehen hundert Chymischen Processen. Frankfurt, 1682. Becher, Johann Joachim. Tripus hermeticus fatidicus, pandens oracular chymica. Frankfurt, 1689. Becher, Johann Joachim. Chymischer Rosen=Garten, Samt einer Vorrede und kurtz gefassten Lebens=Beschreibung Herrn D. Bechers. Ed. Friedrich Roth-Scholtzen. Nuremberg, 1717. Blumenberg, Hans. “Curiosity is Enrolled in the Catalogue of Vices.” The Legitimacy of the Modern Age. Trans. R.M. Wallace. Cambridge: MIT Press, 1983. 309-25. Daston, Lorraine. Eine kurze Geschichte der wissenschaftlichen Aufmerksamkeit. Munich: C.F. von Siemens Stiftung, 2000. Hartung, Gerald. “Johann Joachim Becher.” Die Philosophie des 17. Jahrhunderts. (= Grundriss der Geschichte der Philosophie). Vol. 4.1. Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Basel: Schwabe, 2001. 190-99. Hartung, Gerald. “Johann Joachim Becher oder: Die Projekte und Konzepte eines närrischen Gelehrten.” Zwischen Narretei und Weisheit. Biographische Skizzen und Konturen alter Gelehrsamkeit. Ed. Gerald Hartung and Wolf Peter Klein. Hildesheim: Olms, 1997. 262-87. Klein, Wolf Peter: “Franciscus Mercurius van Helmont.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Vol. 4.1. Basel: Schwabe, 2001. 48-53. Kopp, Hermann. Die Alchemie in älterer und neuerer Zeit. Ein Beitrag zur Kulturgeschichte. Hildesheim: Olms, 1962 [facsimile of the edition Heidelberg, 1886]. Leinkauf, Thomas. “Athanasius Kircher.” Die Philosophie des 17. Jahrhunderts. (= Grundriss der Geschichte der Philosophie). Vol. 4.1. Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Basel: Schwabe, 2001. 269-85.
The “Chymistry Laboratory”
221
Lippmann, Edmund Oskar von. Entstehung und Ausbreitung der Alchemie. 3 vols. Berlin: Springer, 1919. Meier-Oeser, Stephan. “Roger Bacon oder Doctor mirabilis und die Macht der Wahrheit.” Zwischen Narretei und Weisheit. Biographische Skizzen und Konturen alter Gelehrsamkeit. Ed. Gerald Hartung and Wolf Peter Klein. Hildesheim: Olms, 1997. 95-124. Meier-Oeser, Stephan. “Einleitung zu § 1. Hermetisch-platonische Naturphilosophie.” Die Philosophie des 17. Jahrhunderts. (= Grundriss der Geschichte der Philosophie). Vol. 4.1. Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Basel: Schwabe, 2001. 7-18. Molland, A. George. “Roger Bacon as Magician.” Traditio 30 (1974): 445-460. Partington, James Riddick. A History of Chemistry. Vol. 2. London: Macmillan, 1961. Read, John. Prelude to Chemistry. An Outline of Alchemy, its Literature and Relationships. London: G. Bell and Sons, Ltd., 1939. Schmidt-Biggemann, Wilhelm. Philosophia perennis. Historische Umrisse abendländischer Spiritualität in Antike, Mittelalter und Früher Neuzeit. Frankfurt a.M.: Suhrkamp, 1998. Schütt, Hans-Werner. Auf der Suche nach dem Stein der Weisen. Die Geschichte der Alchemie. Munich: C.H. Beck, 2000. Shapin, Steven and Simon Schaffer. Leviathan and the Air-Pump. Hobbes, Boyle, and the Experimental Life. Princeton: Princeton University Press, 1985. Smith, Pamela H. “Chemistry and Commerce. Johann Joachim Becher and the Court of the Elector in Munich.” Johann Joachim Becher (1635-1682). Ed. Gotthardt Frühsorge and Gerhard F. Strasser. Wiesbaden: Harrassowitz, 1993. 143-58. Snelders, Henricus A.M. “Johann Joachim Becher und sein Gold-aus-Sand-Projekt.” Johann Joachim Becher (1635-1682). Ed. Gotthardt Frühsorge and Gerhard F. Strasser. Wiesbaden: Harrassowitz, 1993. 103-14. Stahl, Georg Ernst. Bedencken, Erinnerung und Erlaeuterung über Doctor J. Bechers Natur-Kuendigung der Metallen. Frankfurt and Leipzig, 1723. Stahl, Georg Ernst. Fundamenta Chymica Dogmatico Rationalis et Experimentalis. Nuremberg, 1732. Ströker, Elisabeth. Denkwege der Chemie. Elemente ihrer Wissenschaftstheorie. Freiburg and Munich: Alber, 1967. Ströker, Elisabeth. Theoriewandel in der Wissenschaftsgeschichte. Chemie im 18. Jahrhundert. Frankfurt a.M.: Vittorio Klostermann, 1982. Thorndike, Lynn. A History of Magic and Experimental Science. 8 Vols. New York: Columbia University Press, 1964. Wood, Paul B. “Die philosophische Ausrichtung der Royal Society.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Jean-Pierre Schobinger. Vol. 3.2. Basel: Schwabe, 1988. 391-94. Wood, Paul B. “Robert Boyle.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Jean-Pierre Schobinger. Vol. 3.2. Basel: Schwabe, 1988. 395-403. Yates, Frances A. The Occult Philosophy in the Elisabethan Age. London and Boston: Routledge, 1979.
GERHARD WIESENFELDT
The Order of Knowledge, of Instruments, and of Leiden University, ca. 1700
Judging by accounts of British travelers of the late seventeenth century describing their visits to Leiden University, the architecture of the academy did not seem to represent the high status of its scholarship. In 1691, the Scottish student John Clerk, commenting on the academy building, remarked: “The whole building is by half not so good as an ordinary countrey church.”1 The relatively small number and size of the University buildings was frequently mentioned and contrasted with its English counterparts. In 1691 William Carr remarked: “But as to their Colleges, there are but two, and very small, not to be compared with the smallest Halls in Oxford.”2 Indeed, most of the University buildings were not originally intended for academic use; many of them were former monastery buildings and Catholic churches that stood vacant when Leiden University was founded in 1575, providing provisional accommodation for the teaching activity, and still partly in use today. Over time, additional buildings were acquired to redress the urgent lack of space.3 Besides the large institutions already established in the sixteenth century – the ana1
2
3
John Clerk. Memoirs of the Life of Sir John Clerk of Penicuik. Extracted by Himself from His Own Journals 1676-1755. Ed. John M. Gray. Edinburgh: T.&A., 1892. 13. William Carr. An Accurate Description of the United Netherlands, And of the Most Considerable Parts of Germany, Sweden, and Denmark. London, 1691. 10. Almost the same judgment can be found in William Montague. The Delights of Holland. Or, A Three Months Travel about that and the other Provinces with Observations and Reflections on their Trade, Wealth, Strength, Beauty, Policy, etc. Together with a Catalogue of the Rarities in the Anatomical School at Leiden. London, 1696. 96. On the architectural history of Leiden University, cf. the detailed study by Theodor H. Lunsingh Scheurleer, Cornelia W. Fock, and Albert J. van Dissel. Het Rapenburg. Geschiedenis van eene Leidsche gracht. 15 vols. Leiden: Universiteit Leiden, 1986-1992.
The Order of Knowledge, of Instruments, and of Leiden University
223
tomical theater, the University Library (including fencing and dancing halls on the ground floor), the botanical garden, as well as the Walloon and Staten college for the theology students – a further series of scientific institutions were set up in the seventeenth century – the observatory on the roof of the academy in 1632, as well as the chemical laboratory in 1669 and the physical laboratory in 1675, both at the edge of the botanical garden.4 As a result, an architectural constellation of academic buildings appeared on the Leiden Rapenburg that clearly did not correspond to the expectations of foreign travelers, but which nevertheless played a key role in the University’s self-image as the center of European scholarship. This self-image found its expression in a syllogism, well known during the seventeenth century, which only half in jest described the Rapenburg as the most beautiful street in the world. Even if the individual academic institutions were sometimes extremely small, hardly providing sufficient space for students, they were still assigned an important position within the University, as were the subjects represented by each institution. The existence of this multitude of academic institutions – as small as they were – still reflected the wealth of the scholarship at the University. For the new sciences that had not yet become established in the University curriculum, they also served as proof of a certain standing. In the course of the seventeenth century, the University curators supported such an interpretation of the institutions by assigning one or in the case of the Staten College two professors, and granted priority to these professorships.5 At the same time, each institution stood for a particular form of learning and adopted independent practices. It thus distinguished itself (or had to) from the types of scholarship practiced at other institutions, in order to legitimate its autonomy. These practices and – in case of natural philosophy – the 4
5
On the observatory, cf. Willem Bijleveld, ed. De Leidse Sterrewacht. Vier eeuwen wacht bij dag en bij nacht. Zwolle: Uitgeverij Waanders/De Kler, 1983; on the chemical and physical laboratories, cf. Gerhard Wiesenfeldt. Leerer Raum in Minervas Haus. Experimentelle Naturlehre an der Universität Leiden, 1675-1715. Amsterdam, Diepholz, and Berlin: Verlag für Geschichte der Naturwissenschaften und der Technik, 2002; Willem P. Jorissen. Het chemisch (thans anorganisch chemisch) laboratorium der universiteit te Leiden van 1859-1909 en de chemische laboratoria dier universiteit vóór dat tijdvak en hen, die er in doceerden. Leiden: A.W. Sijthoff’s, 1909. 9ff. This granting of priorities becomes particularly clear in the argument over the chair of chemistry in the 1690s. The medical faculty wanted to reserve this chair for theoretical medicine and make the chemical laboratory part of the botany department. The curators rejected this, referring to the significance of the laboratory. Cf. Wiesenfeldt. Leerer Raum. 210ff.
224
Gerhard Wiesenfeldt
instruments linked to them established new forms of learning and demarcated it from what was represented elsewhere in the university. It would be misleading to assume that this process of distinguishing from other institutions had already been resolved before the establishment of the particular sites. In many cases, this only occurred in subsequent years, but it then played a defining role for the contents and functions of the subject at the University. In this text I want to investigate this process of defining the sciences through the practice of their representation by considering the physical theater. I will thus focus on the way it was delineated from its competing and neighboring institutions: the anatomical theater, the observatory, and the chemical laboratory. Here, I will argue that the distribution of instruments in the sites of knowledge led to decisions that not only concerned the structure of the individual sciences, but also the order of the University as a whole. The Physical Theater as a Site of Independent Natural Philosophy in Leiden The establishment of the physical laboratory in January 1675 was influenced by two opposing factors. As the most recent independent University institution it assimilated the tradition of academic representation mentioned above. In a general European context, it stood within the framework of a development in which the medical and scientific teaching in the second half of the seventeenth century and the early eighteenth century was undergoing a comprehensive transformation resulting from the introduction of experimental collegia and practical instruction.6 The specific circumstances of its founding, however, differed from other institutions at Leiden University. At this institution, it was not primarily a matter of either beginning experimental lectures or establishing the subject at the University. In most scientific fields, the introduction of experiments and practical demonstrations in Leiden had occurred rather casually, without a special institution. Courses with chemical experiments had already been carried out a few decades before the establishment of the chemical laboratory in 1669, partly by local pharmacists, partly by independent lectors, but also by medical professors such as Franciscus de le Boë Sylvius.7 The founding of the laboratory thus con6 7
On natural philosophy in university teaching, cf. ibid. 277ff. Cf. Jorissen. Het chemisch laboratorium; on Sylvius in particular, cf. also Harm Beukers. “Het Laboratorium van Sylvius.” Tijdschrift voor de Geschiedenis van de Geneeskunde, Natuurwetenschappen, Wiskunde en Techniek 3 (1980): 28-36.
The Order of Knowledge, of Instruments, and of Leiden University
225
tributed primarily to the institutionalization of chemistry as an autonomous teaching subject.8 Corresponding to the case of chemistry, there were also private lectures for experimental natural philosophy which were started by the philosophy professor Wolferd Senguerd almost at the same time as the founding of the physical laboratory, but independently of this. These lectures only went on record within the University in 1678 when the curators granted Senguerd a higher salary due to the “large costs of carrying out his experiments that he has occasionally met from his own private purse.”9 Although he kept receiving considerable financial support by the University, Senguerd’s private lectures remained institutionally entirely separate from the physical laboratory until 1705. For Burchard de Volder, Senguerd’s departmental colleague and the initiator of the physical laboratory, it was therefore less a matter of carrying out experiments than of securing the libertas philosophandi, the freedom of academic teaching and study. Since the mid seventeenth century, there had been regular disputes over the philosophy of René Descartes in Leiden as well as in other universities in the Netherlands.10 This conflict particularly concerned the connection between theology and philosophy, a relationship that seemed to have become problematic by Cartesian rationalism. The problem had less to do with the fact that Cartesianism might lead philosophy to demand too great an independence, no longer wanting to assume its traditional role as the handmaiden of theology. Instead, within theology Abraham Heidanus and others aimed at making rationalist philosophy the basis of theology.11 In an attempt to defuse the conflict, the Dutch Grand Pensionary Johan de Witt set out a compromise in 1656, which led to a separation between philosophy and theology. The “Order against 8
9
10
11
Cf. Wiesenfeldt. Leerer Raum. 190ff. For a comprehensive view, cf. Christoph Meinel. “Die Chemie an den Universitäten des 18. Jahrhunderts – Institutionalisierungsstufen und konzeptioneller Wandel.” Academia Analecta. Mededelingen van de Koninklijke Academie voor Wetenschappen, Letteren en Schone Kunsten van België. Klasse der Wetenschappen 48 (1986): 37-57. “De groot oncosten die hem tot het doen van syne Experimenten uyt syn particuliere beurse van tyt tot tyt heeft gedragen.” Leiden University Library, Archief van Curatoren der Leidsche Universiteit, 1574-1815, Ms. 27, fol. 83. On this, cf. Theo Verbeek. Descartes and the Dutch. Early Reactions to Cartesian Philosophy, 1637-1650. Carbondale and Edwardsville: Southern Illinois University Press, 1992; Thomas A. McGahagan. Cartesianism in the Netherlands, 1639-1676. The New Science and the Counter-Reformation. Ph.D. Pittsburgh, 1976; C. Louise Thijssen-Schoutte. Nederlands Cartesianisme. Amsterdam: N.V. Noord-Hollandsche Uitgevers, 1954. Cf. McGahagan. Cartesianism in the Netherlands. 296ff.
226
Gerhard Wiesenfeldt
the mingling of theology with philosophy and the misuse of the freedom of philosophizing to the detriment of the Scriptures” decreed by the States of Holland stated that “all faculties and sciences, including theology and philosophy, have their own set limits within which they must be developed . . . without interfering with the other.”12 While in Leiden both subjects were now to be taught independently for the sake of peace at the University, Cartesian philosophy was no longer to be treated at all – neither approvingly nor disapprovingly. The fact that de Witt, who was a convinced Cartesian, saw himself forced to such a drastic measure demonstrates the level of concern that had emerged in church as well as state circles concerning the conflict at the University. This is not least to be seen against the backdrop of the so-called Remonstrant debate where, in 1618, a conflict between the Leiden theology professors Jacobus Arminius and Franciscus Gomarus over the doctrine of predestination escalated, brought the Netherlands to the verge of civil war, and led among other things to the execution of de Witt’s predecessor, Johan van Oldenbarnevelt.13 While de Witt’s compromise initially released the tensions at the University, the situation deteriorated after 1672 when a state crisis brought about by the English and French invasion culminated in the murder of the Grand Pensionary and the installation of William III of Orange as Stadholder, as well as the rise of the Dutch conservatives. Particularly in church synods, there were vocal demands for the expulsion of theologians and philosophers who supported the ‘French’ philosophy of Descartes. Within the University, these demands found a certain resonance, partly due to serious problems concerning the discipline of students and lecturers that especially affected conservative theologians and philosophers. In 1674, the academic senate had to forbid “the practice of various sorts of tumultuousness, shouting, heckling, throwing of beans and refuse, blaring trumpets as well as other forms of abuse and foolishness” that were causing disruption in lectures.14 As it became clear that the University curators had decided upon hard measures against the Carte12
13 14
The “Ordre jegens de vermeninge van de Theologie met de Philosophie ende het misbruyck van de vryheyt int philosopheren tot nadeel van de Schrifture” is reproduced in Philip C. Molhuysen. Bronnen tot de geschiedenis der Leidsche Universiteit. 7 vols. Den Haag: Nijhoff, 1913-1924. Vol. 3, 55*-58*. Cf. the detailed study in Wiesenfeldt. Leerer Raum. 44ff.; Adriaan C. de Hoog. Some Currents of Thought in Dutch Natural Philosophy, 1675-1720. Ph.D. Oxford, 1974. 47ff. Cf. Jonathan Israel. The Dutch Republic. Its Rise, Greatness, and Fall 1477-1806. Oxford: Clarendon Press, 1995. 361ff., 421ff. Molhuysen. Bronnen. Vol. 3, 280f. Cf. Thijssen-Schoutte. Nederlands Cartesianisme. 276ff.
The Order of Knowledge, of Instruments, and of Leiden University
227
sians, de Volder with the Cartesian-minded theologians Heidanus and Christoph Wittichius strove for a new line of compromise. This contained two basic elements. Firstly, they agreed with the conservative professors at the University that the disturbances of lectures were not acceptable. However, they believed that one should not punish Cartesianism as a whole for this, but only the people directly responsible for the disturbances. They stressed that Descartes himself had always striven for peace, and that his principle of the separation of body and mind ensured precisely the necessary division between philosophy and theology. Secondly, they showed that Cartesianism had come to be valued by the most varied governments, as one could see at the Royal Society, which was established on a Cartesian basis.15 The three professors would have been aware of the fact that the dividing line that they were formulating was no longer one between theology and philosophy, but between metaphysics and natural philosophy, just as the reference to the Cartesianism of the Royal Society must have seemed rather strained to their contemporaries. Both formulations however point to the future of philosophy as independent of theology. Following this attempt at mediation, de Volder undertook a trip to England, visited Isaac Newton in Cambridge and made contact with members of the Royal Society.16 Immediately after his return he made a request to the curators asking them to put a place at his disposal where “according to the model of foreign academies and illustrious schools . . . by experiments the truth and certainty of the theses and teachings that students are taught in physical science should be shown.”17 Without anyone mentioning the Cartesianism debate, it was nevertheless clear that this was largely a matter of giving philosophy a site of representation where it could be prominently represented within the University without the threat of disputes about the drawing of borders between philosophy and theology. The curators understood this intention very well and supported de Volder enthusiastically by not only granting his request, but provided around 3,500 guilders in 1675 alone and thus roughly an eighth of the University budget for the purchase of a building and the acquisition of the necessary instruments, despite the fact that the University was undergoing financial difficulties at the 15
16 17
Jean Le Clerc. “Éloge de feu Mr. de Volder, Professeur en Philosophie et aux Mathématiques, dans l’Académie de Leide.” Bibliothèque choisie 18 (1709): 357; cf. Wiesenfeldt. Leerer Raum. 58ff. Cf. de Hoog. Some Currents of Thought. 144f. Molhuysen. Bronnen. Vol. 3, 298.
228
Gerhard Wiesenfeldt
time.18 Subsequently, they promptly introduced drastic measures against the Cartesians at the University, though no longer banning Cartesian philosophy as a whole, but explicitly Cartesian metaphysics.19 The implicit permission to deal with Cartesian natural philosophy has to be seen in relation to the physical laboratory, which now represented a free and unproblematic form of philosophy. As for the rest, de Volder rigorously ignored the ban on advocating Cartesian metaphysics, though only in private lectures and in the auditorium majorum of the University. The physical laboratory exclusively served experimental physical science, which de Volder orientated as far as possible towards the program of the Royal Society and in particular to Robert Boyle’s New experiments physico-mechanical. At the center of the laboratory was an air-pump that had been built by the Leiden instrument maker Samuel van Musschenbroek upon a design that de Volder had brought back from London. Without adopting Robert Boyle’s abundant rhetoric, de Volder made it clear by this orientation that he had recognized the irenic potential of an experimental natural philosophy with Boyle’s stamp.20 Demarcation between Chemistry and Physical Science The setting up of the physical laboratory, however, did not proceed entirely without difficulties owing to a strong protest made to the curators by the professor of chemistry, Carel de Maets, who felt that this would create an institution that would be a competitor to the chemical laboratory. Chemistry had recently gone to great lengths to establish itself as an independent department within the philosophical faculty after having long played a marginal role as an auxiliary discipline between medicine and botany with the production of pharmaceutics. Only in 1658, with the appointment of the iatro-chemist Franciscus de le Boë Sylvius as professor of theoretical medicine, did it gain independent significance as a subject that was able to elucidate the foundations of bodily processes. Sylvius was also the driving force behind plans to establish a separate chair of chemistry, which, however, could not be realized after attempts to appoint Boyle failed due to his wealth.21 Instead, in 1669, Sylvius 18 19 20
21
Cf. Wiesenfeldt. Leerer Raum. 61ff. Molhuysen. Bronnen. Vol. 3, 318. On Boyle’s program, cf. Steven Shapin and Simon Schaffer. Leviathan and the Air-Pump. Hobbes, Boyle, and the Experimental Life. Princeton: Princeton University Press, 1985. Cf. Molhuysen. Bronnen. Vol. 3, 196.
The Order of Knowledge, of Instruments, and of Leiden University
229
managed to set up the laboratory with de Maets appointed as director. The latter was then quickly appointed extraordinary professor and in 1672 full professor. At the instigation of de Maets and Sylvius, who was rector during this year, he was assigned to the philosophical faculty against their resolute will. This appointment was also behind his protest against the physical laboratory. According to his arguments, there was already a place for experiments in the philosophical faculty, that is the chemical laboratory assigned to him. In addition, he could carry out the same experiments as de Volder significantly cheaper. While the financial argument did not play a role for the curators, the competition between the two places posed a certain dilemma since they wanted to support the physical laboratory, though without passing over the chemical laboratory. After a failed attempt at mediation, they decided that both professors in their respective institutions should be granted the right to carry out proofs in experimental philosophy, and also increased the annual budget of the chemical laboratory.22 The following period witnessed regular conflicts between both sites for the representation of experimental science, even after the chair of chemistry was transferred back to the medical faculty in 1678, since it had “more affinity and association with the medical faculty than with philosophy” according to the curators.23 The dilemma in these conflicts was that both chemistry and natural philosophy claimed to describe the fundamental principles of nature with their experiments, particularly in relation to the human body. The areas of investigation of both subjects constantly overlapped to the extent that de Maets’ successor, Jacob le Mort, claimed that experimental natural philosophy was merely a branch of chemistry.24 Furthermore, representatives of both fields claimed that they stood in the tradition of Boyle. In view of this overlap, one might wonder why both subjects developed separately. Here, the manner in which both places were equipped with instruments was essential, since independently of the object of investigation and the aim of research, the chemical laboratory remained the place where experiments that required smelting stoves were performed, while the physical laboratory was the place for experiments using the air-pump and barometer. The demarcation between both places 22 23 24
Cf. Wiesenfeldt. Leerer Raum. 199ff. Leiden University Library, Archief van Curatoren der Leidsche Universiteit, 15741815, Ms. 27, fol. 111. Cf. Wiesenfeldt. Leerer Raum. 210ff. Cf. Christopher Love Morley. Collectanea chymica Leydensia. Jena, 1696.
230
Gerhard Wiesenfeldt
corresponded as little to a logic of research as the separation between the physical laboratory and anatomical theater, where experiments on dogs were carried out for the investigation of life processes in both places.25 Empiricism as Unifying Element Although instruments served as a means of demarcation between the different sites of experimental science, they also had a unifying function that was likewise negotiated in the curators’ meetings in which de Volder and de Maets argued about the assignation of experimental physical science. After they had presented their opposing positions, the botanist Arnold Seyen, then rector of the University, began to speak. Seyen was clearly concerned that the curators would reject the claims of both professors on account of their argument, and as a result gave a long speech in which he described his experience of giving practical lectures in botany. He stressed how much more difficult such a class is compared to a purely theoretical one, how much more work he would invest in the preparation of lectures, and how much more the students would learn in them, and finally how much satisfaction he was given by the fulfillment of this difficult task.26 The logic of this speech not only aims at defusing the argument between de Maets and de Volder, but also relates to the growing self-understanding of the medical faculty, in which its sites of representation – the anatomical theater and the botanical garden – were the expression of an ideal of knowledge that was no longer based on the model of the philological sciences still prevalent at the University, but on direct observation and the experience derived from this.27 Because of de Maets and de Volder, this ideal had now spread to the philosophical faculty, against whose speculative orientation the medical faculty had repeatedly spoken out during the debate on Cartesian25
26 27
Cf. Gerrit A. Lindeboom. “Dog and Frog. Physiological Experiments at Leiden during the Seventeenth Century.” Leiden University in the Seventeenth Century. An Exchange of Learning. Ed. Theodor H. Lunsigh Scheurleer and Guillaume H.M. Posthumus Meyjes. Leiden: Brill, 1975. 279-293; Wiesenfeldt. Leerer Raum. 117ff. Molhuysen. Bronnen. Vol. 3, 299. On the anatomical theater, cf. Theodor H. Lunsingh Scheurleer. “Un amphithéâtre d’anatomie moralisée.” Leiden University in the Seventeenth Century. An Exchange of Learning. Ed. Theodor H. Lunsingh Scheurleer and Guillaume H.M. Posthumus Meyjes. Leiden: Brill, 1975. 217-277; Jan Rupp. “Matters of Life and Death. The Social and Cultural Conditions of the Rise of Anatomical Theatres, with Special Reference to Seventeenth-Century Holland.” History of Science 28 (1990): 263-87.
The Order of Knowledge, of Instruments, and of Leiden University
231
ism.28 The transformation that then took place within the philosophical faculty was as swift as it was comprehensive. In 1671 an independent core faculty had been created whose members claimed for themselves the exclusive right to confer doctorates, and in this way aimed particularly at undermining the dominance of the philologists within the faculty.29 After 1676 this core faculty included – in addition to de Maets and de Volder – the mathematician Christian Melder, who was also director of the observatory, and the philosopher Senguerd, also an active experimentalist. The self-understanding of natural philosophers and physicians is revealed precisely in cases of conflict with other faculties – as scholars who support their work with empirical facts as well as, in particular, with instruments. From Symbolic Site to Symbolic Instrument At the beginning of the eighteenth century, for empirical science, the significance of the sites of representation in Leiden became less important for the self-understanding of these sites than the instruments contained in them. After de Volder was relieved from teaching duties in 1705, Senguerd took over the directorship of the physical laboratory, but continued to give most of his experimental lectures in his home. After 1724, Senguerd’s successor, Willem Jacob ’s Gravesande, also relied more on his private instrument cabinet than the now completely outdated cabinet in the physical laboratory. In the same way, de Maets’ successor, Jacob le Mort, used, besides the chemical laboratory, his own private laboratory as well as the anatomical theater for his lectures after his appointment in 1702. For some time after 1707, the observatory was hardly used.30 Gradually, the strictly content-based separation of the different institutions was increasingly undermined. This became particularly clear in 1718 when Herman Boerhaave took over the chair of chemistry. Boerhaave had been the student of the natural philosophers Senguerd and de Volder, but was also, at least since his appointment to the chair of botany in 1709, well established in the medical faculty. In a 28 29
30
Cf. Molhuysen. Bronnen. Vol. 3, 107. “Acta et Decreta Facultatis Philosophicae.” Leiden University Library, Archieven van Senaat en Faculteiten der Leidsche Universiteit (1575-1877): Ms. 462, fol. 1. On the role of philology before this, cf. Anthony Grafton. “Civic Humanism and Scientific Scholarship at Leiden.” The University and the City. From Medieval Origins to the Present. Ed. Thomas Bender. Oxford: Oxford University Press, 1988. 59-78. Cf. Wiesenfeldt. Leerer Raum. 95, 219ff.
232
Gerhard Wiesenfeldt
series of programmatic orations he carefully strove for a division between chemistry and natural philosophy, in which both would examine the foundations of natural processes, though chemistry would be responsible for the basic operations and natural philosophy for the basic principles of nature.31 However, immediately after assuming the chair, he began to undermine this division by placing an air-pump in the chemical laboratory, experimented with it, and, furthermore, started a collaboration with the Amsterdam instrument maker Daniel Gabriel Fahrenheit on matters that in Leiden had previously been assigned to experimental philosophy.32 At the same time, the significance of the instruments for the different subjects remained unbroken to the extent that around 1720 the instruments took on an emblematic function in the representation of the corresponding faculty in academic publications,33 which is particularly astonishing for chemistry within the medical faculty. Consequently, in experimental philosophy, Leiden was no longer pre-eminently the city of the physical laboratory, but rather the town where the Senguerdian airpump was produced, which had now become the standard instrument of experimental philosophy.34 If there was still a central site for experiments in Leiden, then it was the workshop of the instrument-making family van Musschenbroek, situated directly next to the University.35 Practically all the Leiden professors active in the experimental field were supplied with air-pumps, glass instruments, telescopes, microscopes, as well as mechanical and surgical instruments from this workshop. This place, also, was not without a high symbolic value, since for the professors of other universities, the purchase of Musschenbroek instruments and their use in their own experimental collegia was the proof of a direct orientation towards the model of the leading academy in Europe.36
31 32
33 34 35
36
Herman Boerhaave. Boerhaave’s Orations. Ed. and trans. Elze Kegel-Brinkgreve and Antonie M. Luyendijk-Elshout. Leiden: Brill, 1983. Cf. Gerrit A. Lindeboom. Herman Boerhaave. The Man and his Work. Leiden and London: Methuen, 1968. 340-44; Daniel Gabriel Fahrenheit. Fahrenheit’s Letters to Leibniz and Boerhaave. Ed. and trans. Pieter van der Star. Amsterdam: Rodopi, 1983. This is the case in, for instance, the disputation texts published by Daniel Goetval. Cf. Anne C. van Helden. “The Age of the Air-Pump.” Tractrix 3 (1991): 149-72. On the history of the Musschenbroek workshop, cf. Peter de Clercq. At the Sign of the Oriental Lamp. The Musschenbroek Workshop in Leiden, 1660-1750. Rotterdam: Erasmus Publications, 1997. Cf. Peter de Clercq. “Exporting Scientific Instruments around 1700. The Musschenbroek Documents at Marburg.” Tractrix 3 (1991): 79-120.
The Order of Knowledge, of Instruments, and of Leiden University
233
Thus, even from an external perspective, the experimental sciences in Leiden had become emancipated from the sites of their representation. This however raises the question whether the unifying element of the sciences represented in these sites is not based on the reference to an artisanal ideal which requires an instrument, but only encompasses a part of what was understood by an experiment. At no point the actual experimental labour performed in the sites of representation, which could have led to a thematic figuration of the disciplines, possessed an essential significance for the self-understanding of the disciplines represented.
WORKS CITED Beukers, Harm. “Het Laboratorium van Sylvius.” Tijdschrift voor de Geschiedenis van de Geneeskunde, Natuurwetenschappen, Wiskunde en Techniek 3 (1980): 28-36. Bijleveld, Willem, ed. De Leidse Sterrewacht. Vier eeuwen wacht bij dag en bij nacht. Zwolle: Uitgeverij Waanders/De Kler, 1983. Boerhaave, Herman. Boerhaave’s Orations. Ed. and trans. Elze Kegel-Brinkgreve and Antonie M. Luyendijk-Elshout. Leiden: Brill, 1983. Carr, William. An Accurate Description of the United Netherlands, And of the Most Considerable Parts of Germany, Sweden, and Denmark. London, 1691. Clercq, Peter de. “Exporting Scientific Instruments around 1700. The Musschenbroek Documents at Marburg.” Tractrix 3 (1991): 79-120. Clercq, Peter de. At the Sign of the Oriental Lamp. The Musschenbroek Workshop in Leiden, 1660-1750. Rotterdam: Erasmus Publications, 1997. Clerk, John. Memoirs of the Life of Sir John Clerk of Penicuik. Extracted by Himself from His Own Journals 1676-1755. Ed. John M. Gray. Edinburgh: T.&A., 1892. Fahrenheit, Daniel Gabriel. Fahrenheit’s Letters to Leibniz and Boerhaave. Ed. and trans. Pieter van der Star. Amsterdam: Rodopi, 1983. Grafton, Anthony. “Civic Humanism and Scientific Scholarship at Leiden.” The University and the City. From Medieval Origins to the Present. Ed. Thomas Bender. Oxford: Oxford University Press, 1988. 59-78. Helden, Anne C. van. “The Age of the Air-Pump.” Tractrix 3 (1991): 149-72. Hoog, Adriaan C. de. Some Currents of Thought in Dutch Natural Philosophy, 16751720. Ph.D. Oxford, 1974. Israel, Jonathan. The Dutch Republic. Its Rise, Greatness, and Fall 1477-1806. Oxford: Clarendon Press, 1995. Jorissen, Willem P. Het chemisch (thans anorganisch chemisch) laboratorium der universiteit te Leiden van 1859-1909 en de chemische laboratoria dier universiteit vóór dat tijdvak en hen, die er in doceerden. Leiden: A.W. Sijthoff’s, 1909. Le Clerc, Jean. “Éloge de feu Mr. de Volder, Professeur en Philosophie et aux Mathématiques, dans l’Académie de Leide.” Bibliothèque choisie 18 (1709): 346-401. Lindeboom, Gerrit A. Herman Boerhaave. The Man and his Work. Leiden and London: Methuen, 1968. Lindeboom, Gerrit A. “Dog and Frog. Physiological Experiments at Leiden during the Seventeenth Century.” Leiden University in the Seventeenth Century. An Exchange
234
Gerhard Wiesenfeldt
of Learning. Ed. Theodor H. Lunsigh Scheurleer and Guillaume H.M. Posthumus Meyjes. Leiden: Brill, 1975. 279-93. Lunsingh Scheurleer, Theodor H. “Un amphithéâtre d’anatomie moralisée.” Leiden University in the Seventeenth Century. An Exchange of Learning. Ed. idem and Guillaume H.M. Posthumus Meyjes. Leiden: Brill, 1975. 217-77. Lunsingh Scheurleer, Theodor H., Cornelia W. Fock, and Albert J. van Dissel. Het Rapenburg. Geschiedenis van eene Leidsche gracht. 15 vols. Leiden: Universiteit Leiden, 1986-1992. McGahagan, Thomas A. Cartesianism in the Netherlands, 1639-1676. The New Science and the Counter-Reformation. Ph.D. Pittsburgh, 1976. Meinel, Christoph. “Die Chemie an den Universitäten des 18. Jahrhunderts – Institutionalisierungsstufen und konzeptioneller Wandel.” Academia Analecta. Mededelingen van de Koninklijke Academie voor Wetenschappen, Letteren en Schone Kunsten van België. Klasse der Wetenschappen 48 (1986): 37-57. Molhuysen, Philip C. Bronnen tot de geschiedenis der Leidsche Universiteit. 7 vols. The Hague: Nijhoff, 1913-1924. Montague, William. The Delights of Holland. Or, A Three Months Travel about that and the other Provinces with Observations and Reflections on their Trade, Wealth, Strength, Beauty, Policy, etc. Together with a Catalogue of the Rarities in the Anatomical School at Leiden. London, 1696. Morley, Christopher Love. Collectanea chymica Leydensia. Jena, 1696. Rupp, Jan. “Matters of Life and Death. The Social and Cultural Conditions of the Rise of Anatomical Theatres, with Special Reference to Seventeenth-Century Holland.” History of Science 28 (1990): 263-87. Shapin, Steven and Simon Schaffer. Leviathan and the Air-Pump. Hobbes, Boyle, and the Experimental Life. Princeton: Princeton University Press, 1985. Thijssen-Schoutte, C. Louise. Nederlands Cartesianisme. Amsterdam: N.V. NoordHollandsche Uitgevers, 1954. Verbeek, Theo. Descartes and the Dutch. Early Reactions to Cartesian Philosophy, 1637-1650. Carbondale and Edwardsville: Southern Illinois University Press, 1992. Wiesenfeldt, Gerhard. Leerer Raum in Minervas Haus. Experimentelle Naturlehre an der Universität Leiden, 1675-1715. Amsterdam, Diepholz, and Berlin: Verlag für Geschichte der Naturwissenschaften und der Technik, 2002.
ANGELA MAYER-DEUTSCH
The Ideal Musaeum Kircherianum and the Ignatian Exercitia spiritualia
In Miguel Guerrero’s frontispiece to Francisco de Florencia’s1 Historia de la Provincia de la Compania de Jesus da Nueva-Espania (Mexico 1694) the process of conversion is depicted as an effect of light rays (fig. 1). In the upper part of the picture the three saints acting in it are shown in a triangular composition. There is a pedestal in the center of the lower part into which the book title and a much-distorted map of Central America are inscribed. The map shows the “terra incognita” of those regions that have not yet been Christianized. The pedestal is flanked by two groups of men. Emanating from the chest of Saint Francisco de Borja,2 which carries the coat of arms of the Society of Jesus widened into a blazing triangle, two groups of rays shoot to his right and his left onto the coats of arms of the two saints positioned underneath, Ignatius of Loyola and Francis Xavier. These coats of arms are surrounded by an aureole and seem to operate like a prismatic mirror (without being depicted as such), as they diffract and at the same time dissect the rays into three smaller rays which are directed towards two groups of Creoles with opulent headdresses awaiting their conversion on their knees. This frontispiece is a variation of Pierre Miotte’s earlier frontispiece to the “optical” foliant Ars Magna Lucis et Umbrae (Rome 1646 and 1671) by Athanasius Kircher. The latter is designed as an allegory of the universal power of light and its effects, and also of the forms of insight it enables. In the case of Florencia, the divine light – and the truth that is associated with it – radiate from one heart to the other. In the case of Kircher the divine light and the divine truth are mediated by the Holy Scripture and by mirrors, whose effect is reinforced by a telescope and a peacock and thus associated with natural 1 2
Florencia was professor of theology and philosophy and procurator of the Jesuit province of New Spain, later Mexico. The Jesuit General Borja inaugurated missions in Florida, Peru, and Mexico.
236
Angela Mayer-Deutsch
Fig. 1: Frontispiece of Francisco de Florencia. Historia de la Provincia de la Compania de Jesus da Nueva-Espania (Mexico, 1694).
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
237
philosophy in general and theories of optics in particular. The frontispiece to Kircher thematizes the connection between natural philosophy and conversion rather indirectly, but it becomes evident once the two engravings are subjected to a comparative viewing. Nevertheless, it will be the point of departure for some reflections on the “active contemplation” of pictures in Kircher’s ideal museum.3 Ways to the Heart or the Accompaniment of the Exercises Kircher was a polymath, a “museum’s author,”4 and a Jesuit. His impact was above all determined by his spiritual concerns as a priest and by an “active contemplation” of various paths leading to God. Philosophical and religious insight went hand in hand here. God’s providence required him “to use the small amount of talent granted to me despite my unworthiness to worship his Holy Name and to advance public welfare.”5 “Active contemplation” was used in Ignatian prayer and in the Exercises with the aim of giving a concrete shape to various locations and events of the life of Jesus, and it was realized by means of a reinforced investment of imagination.6 It required the engagement of the entire person, but especially the heart as the site of spiritual life. The spiritual exercises were practiced within a four week period that involved the renunciation of daily life; the disciple or exercitant was taught to experience a number of highly divergent emotional states which appealed to all senses, but above all to the visual sense. By passing through and overcoming those emotional states, the exercitant was guided away from material things and led towards an immaterial truth and to the recognition of God’s will. Finally, the exercitant would obtain divine grace in the form of a vision, for example. According to Ignatius, man’s pas3
4
5
6
In my opinion it is the ideal museum – rather than the actual collection – that represents Kircher’s general concept of the world, and it is this view of the world that is supposed to be contemplated and finally mediated in the museum. There is a division or redoubling of the author’s function in the museum’s catalogue: Kircher’s machine-operator Giorgio de Sepi is the author of the catalogue while Kircher is the author of the museum itself. Similar divisions of authorship can be found in Kircher’s correspondence. Nikolaus Seng. Selbstbiographie des P. Athanasius Kircher aus der Gesellschaft Jesu. Fulda: Druck und Verlag der Fuldaer Actiendruckerei, 1901. 49. The source of this translation is the Latin autobiography published in 1684 and probably written in the beginning of the 1670s. Although the Exercises are still practiced today, I decided to use the imperfect to make it clear that I am talking about the practice of the seventeenth century.
238
Angela Mayer-Deutsch
sage towards experience, change, and growth was guided by the Holy Spirit; for it was He who let man experience the love and the grace of God by writing or almost engraving it into his open heart. Originally the Exercises were written for the companion or director, who chose certain exercises, passages from the Bible, songs, etc. to guide the exercitant on his way towards divine grace and the Holy Spirit. Similarly to the director of the Exercises, Kircher used to select certain passages from the Bible as gifts to museum’s visitors or within letters to correspondents. These were always connected with the person or his actual situation in life. In order to emphasize the persuasive power of the Catholic world, for instance, he once gave a model of an obelisk to Queen Christina of Sweden after her second visit to the museum. In addition to the model, which had been made for this particular occasion, Christina received an Arabic translation of the Psalms of David including special indications of the passages describing Solomon’s temple and Moses’s tabernacle. Kircher argued that both edifices represented the Church of Rome and at the same time alluded to Christina’s plan of founding an academy in Rome. His practice of guiding thus resembles that of the director of the Exercises, which will be explored in further detail in the following. The museum was located in the Collegium Romanum and was closely connected with its “author” Kircher, who, in the frontispiece of the catalogue, is portrayed as touching his heart with his left hand.7 The human heart was supposed to be ignited by the flame of divine love during the final phase of contemplation in the fourth and last week of the spiritual journey, the point of culmination of the entire Exercises (the contemplation aimed at the attainment of love). The flame of love facilitated the exercitant’s experience of partaking in Christ. It is no coincidence that both Jesuit contemplation and Kircher’s museum revolved around the central motif of the heart: Ignatius himself was a worshipper of the heart of Jesus.8 He was extremely impressed by the legend of the martyr-bishop Ignatius of Antioch, which explains more than the mere fact that he changed his name from Inigo to Ignatius; his obsession with 7
8
Following the common practice of heart burials for members of the reigning Catholic houses as well as for clergymen, Kircher chose as a cemetery for his heart St. Eustachio di Mentorella near Guadagnolo, in the southeast of Rome (near Tivoli). It was buried at the foot of the altar. To be precise, the cult of the heart of Jesus only began between 1673 and 1675 in Paray-le-Monial with the visions of Margaretha Maria Aloque, a Salesian nun. Nonetheless it has its early origins and sources in the spirituality of the Carthusian monastery of the thirteenth and fourteenth century and in the Exercises.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
239
the legend is also reflected in the name of the Society as well as in the stamp of the first seal. One reads that St. Ignatius, despite all the torments he had to endure, never forgot the name of Christ. And so the torturers that tormented him asked why he frequently pronounced that name. He answered, “I cannot stop myself, since it is written into my heart.” After his death the Pagans wanted to verify this, and they ripped his heart out of his corpse and opened it; there, in golden letters, was the name of Jesus Christ written in its center. This sign effected the conversion of numerous Pagans.9
The principal focus of Kircherian natural philosophy was the magnet. Every phenomenon in the world worked according to the principles of attraction and repulsion. God was the central magnet of the universe and acted ‘magnetically.’ His explication of God and the Holy Trinity in reference to an inconspicuous loadstone in Magnes (1641) brought Kircher towards the threshold of heresy. Magnetic demonstrations – which always implied an analogy with the ‘divine magnet’ and its creations – as well as hydraulic and pneumatic demonstrations were some of the earliest spectacles staged in the cubiculum and, from 1651 onwards, also in the official museum. One of the most disputed highlights of those shows was a magnetic sunflower clock, which was made of a sunflower fixed on a disc of cork that floated in a bowl of water and which, in its circular floating motion, followed the direction of the sun. There was a mark on the cork disk that pointed to the dial painted on the edge of the bowl and thus indicated the time. According to Kircher, the attraction of the sun towards which the flower orientated itself was magnetic. And so he called the clock magnetic too. The idea of such a magnetic clock had already been developed among others by Galileo Galilei, who conceived of the cork disk as ‘little earth’ and connected it via terella to the idea of the moving earth.10 The catalogue of the museum contains a description of the magnet as the ‘heart’ of the collection, which is situated in the middle of the book: The magnet – challenge to subtle spirits, miracle of nature, magician of wonders and impenetrable labyrinth of hidden virtue, the heart of the Musaeum Kircherianum – is a stone which is blessed neither with any shine nor any par9 10
Jacobus de Voragine. Legenda aurea. Jena: Diedrichs, 1925. 184. Until 1633 Nicolas Claude Fabri de Peiresc hoped to find Kircher’s support in defense of Galileo, who claimed that the magnetic clock gave evidence to the heliocentric worldview. Kircher, however, decidedly turned his back on the Copernican doctrine after the decree of 1633 and became an acknowledged geocentrist. In 1641 he was the first to dedicate an entire book to a meticulous attack on Kepler’s as well as Gilbert’s magnetic astronomy.
240
Angela Mayer-Deutsch
ticular beauty as far as its formal appearance is concerned. But it is endowed with such unusual and unique virtues that it is beyond comparison with anything else. And the human community does not benefit from anything as much as it benefits from the magnet.11
Keeping Kircher’s gesture on the frontispiece of the catalogue in mind we can say that the heart – as the focus of the Exercises and as a metaphor of the magnetic core of the museum – represented that particular part of man that could be affected. And this process of affectating the heart might have had an equivalent in the individual guided tours that Kircher himself conducted – yet another analogy to the director of the Exercises. Accordingly the Protestant Margrave of Brandenburg, Christian Ernst, was presented with a model of the city of Jerusalem at the time of Christ’s Passion. This model had probably been produced by Juan Battista Villalpando, S.J. and was shown at the entrance of the museum. The second volume of Villalpando’s commentary on the prophecy of Ezekiel (1604)12 included a reconstruction of Solomon’s temple as a prototype of the celestial Jerusalem. Villalpando argues that a complete record of the prophet’s mystical message had to be made manifest as a translation into actual architecture; and apparently such a translation of the entire contemporary city had indeed taken place, at least as a translation into a plaster model. Consequently, the replica of Jerusalem could be interpreted as both a symbol of a ‘correct’ attitude towards vision and a symbol of the Kunstkammer (not only Kircher’s)13, which could generally be conceived of as a ‘translation’ of the house of Solomon. Finally, to the margrave of Brandenburg, it might have been a symbol of the Catholic Church of Rome. Therefore Kircher consciously decided to draw the Protestant nobleman’s attention to the model of Jerusalem 11
12 13
“Magnes subtilium cos ingeniorum, prodigium Naturae, mirabilium thaumaturgus, reconditae virtutis inscrutabilis Labyrinthus, Kircheriani Musaei medulla, lapis est, qui quoad formalem apparentiam nil neque splendoris neque gratae pulchritudinis praerogativa gaudet, cum tamen tam eximia, & singulari virtutis dote praeditus sit, ut nulli lapidum speciei secundus habendus sit, nec ab ullo, quod ad humanum commercium specatat, tantum utilitatis recipitur, quam à Magnete, hic enim solus navigationes dirigit, & ad desideratos, velificantes in erronei maris incognitis semitis, portus deducit . . .” Giorgio De Sepi. Musaeum celeberrimum. Amsterdam, 1678. 18. The spot is to be found in the fourth chapter of the second of three parts of the catalogue, almost in the ‘heart’ of it. Juan Battista Villalpando. Explanationes in Ezechielis et apparatus urbis ac templi Hierosolymitani. Rome, 1604. Before Kircher, Villalpando had dedicated much thought to the details of the measurements of another Biblical architectural design, namely Noah’s ark. Kircher elaborated this idea in an entire book: Athanasius Kircher. Arca Noê . . . Amsterdam, 1675.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
241
in order to emphasize the universal significance of the Catholic Church in contrast to the (‘inferior’) Protestant church. There were two other visitors to the museum whom Kircher guided or ‘accompanied’ in a distinctly different manner: the already mentioned Queen Christina, who had just converted to Catholicism, and then the future secretary of the Royal Society of London for the Improvement of Natural Knowledge, Robert Southwell. Both were interested in seeing the same ‘experiment,’ but their interpretations of the demonstrations in 1656 and in 1661 varied to a great extent. Christina saw a number of magnetic experiments and a “hermetic” demonstration, the so-called “palingenetic” experiment,14 as a highlight of her visit to the museum. In analogy to the resurrection of Christ it was also called the “vegetable phoenix.” But what Christina was actually shown were only the results of the experiment: a tiny plant and its ashes in a phial. She was told that the plant had grown out of its own ashes in the absence of sunlight, which, in the case of Christina, was interpreted as a reference to her own ‘revival.’ The event of her conversion and more generally the triumph of the ‘resurrection’ of the Catholic Church was the major subject of all festivities and celebrations of her arrival in Rome. It was accompanied by the precious gift already mentioned above, the model of the obelisk. Both the obelisk and Christina’s visit were eloquently described and honored in the catalogue more than twenty years after (De Sepi 1678). The model included the following inscription: “Athanasius Kircher S.J. made, named, and consecrated this obelisk, inscribed with arcane characters of the ancient Egyptians for the great Christina.”15 According to the catalogue the inscription was written in thirty-three languages. The fact that Kircher saw Christina as the equivalent of Isis and similarly treated his patron pope Alexander VII as the equivalent of Alexander the Great was related to his attempts to stylize and present the museum not only as a reflection of contemporary Rome but also as the image of a second ‘resurrected’ Alexandria, with himself as a guide in charge of mediation, translation and explanation: a new ‘Hermes (Trismegistos).’ In contrast, Robert Southwell’s information about the “hermetic” experiment was based on textual sources, namely the Mundus Subter14
15
The term “palingenesia” derives from Greek SDOLQJHQHVLD, translated into Latin as “iterata generatio.” Cf. Heinrich Georges. Lateinisch-Deutsches Handwörterbuch. Vol. 2. Hanover, 1892. Col. 1444. Later it was used synonymously with resurrection. “Magnae Christinae, Isidi redivivae, Obeliscum hunc arcanis, Veterum Aegyptiorum notis inscriptum erigit, dicat, consecrat A.K.S.J.” De Sepi. Musaeum. 12.
242
Angela Mayer-Deutsch
raneus, and also on various reports on the Queen’s visit. He had been planning to travel to Rome since 1657 to find out how and if it actually worked. In 1661 he wrote from Rome to Robert Boyle, another founding member of the Royal Society: Father Kircher is my particular friend, and I visit him and his gallery frequently. Certainly he is a person of vast parts, and of great industry. He is likewise one of the most naked and good men that I have seen, and is very easy to communicate whatever he knows; doing it, as it were, by a maxim he has. On the other side he is reputed to be very credulous.16
But he did not see anything of the legendary “vegetable phoenix” because the plant was now dead. According to various reports, the phial had broken due to cold weather. Southwell also reported that the plant had merely been given to Kircher and therefore could not be called his own product. But at least he returned home with the recipe and information about all the ingredients required. Unfortunately there is no record of a successful replication of the experiment. Consequently, Christina’s visit must have been very well prepared and carefully carried out, especially with regard to the rhetoric of conversion. Southwell and Kircher, on the other hand, probably mainly engaged themselves in conversation, since Southwell was neither a good candidate for conversion nor a useful subject for propaganda. Christina interpreted the experiment as the ‘resurrection’ of a plant, analogous to her own conversion. Since the gentlemen of the Royal Society had simply been interested in the possibility of a plant’s growing out of its own ashes when totally secluded from sunlight, they ridiculed Kircher for his interpretation. The Rhetoric of Images and the ‘Emblematic’ Worldview Kircher’s goodness, “nakedness,” and perfect openness, as described by Southwell, were praised over and over again by various people. It is possible that even John Evelyn’s report about the museum in Kircher’s cubiculum, which is generally rather disparaging in its tone, refers to the Jesuit’s impeccable character in a remark on his “dutch patience.” I would argue that this goodness, patience, and readiness for convers(at)ion were indefatigably effective both in the museum and in his work. His aim was not only to (re)gain souls for Catholicism but also to propagate 16
Southwell to Boyle, Rome, March 30, 1661. Robert Boyle. The Works. Vol. 6. Ed. Thomas Birch. Hildesheim: Olms, 1966. 299.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
243
an ‘emblematic’ worldview that was shared by philosophers like Claude Menestrier or Emanuele Tesauro. Their methodological attempts to reconstruct a unity that had been lost or had fallen silent were based on the belief that there was, in principle, a unity of the world. Images played an important role in this conception of the world conceived of as a reflection of God. The ‘emblematic’ idea of nature as an elaborate hieroglyph,17 as a source of miracles and mystery did not lead to a distinction between true and untrue phenomena. This distinction seems to be irrelevant for authors like Kircher, Giambattista Riccioli, Francesco Lana Terzi, or Caspar Schott. For any attempt at disclosure inevitably failed to reach the veiled, divine truth. This truth could not be and was not to be revealed in its entirety. Both in natural philosophy as well as theology it remained a game with a number of hypotheses whose probability was not essential. Therefore eclecticism, fictionalism, and probabilism were the outstanding structural elements of ‘Jesuit natural philosophy’ in the seventeenth century, as we will see in the following reflection on certain engravings. The continuous and conscious use of images as well as “active contemplation” aimed, I would argue, at the consolidation of faith and conversion. This was also true for the training of perception by means of deceptions and confusions that could be observed both in the works and the practices of each of the named Jesuit authors. Such an effect of consolidation was expected from the Exercises, which are essential to every Jesuit, and also from the demonstrations in the museum.18 In a broader sense, both practices were intended to civilize people and offer exercise in piety to visitors from all parts of the world. During that process the image itself was merely regarded as an aid, but not as object of worship. Only the prototype deserved worship – an important concept in the theology of images of that time. The Jesuits specifically learned to use internal and external images as aids in their meditation exercises. This function is evidenced by the generally important role emblems always 17
18
Cf. William B. Ashworth, Jr. “Natural History and the Emblematic World View.” Reappraisals of the Scientific Revolution. Ed. David C. Lindberg and Robert S. Westman. Cambridge: Cambridge University Press, 1990. 303-32 about the development of this worldview. It is derived from six traditions and attributed to sixteenth century natural philosophy and to some Jesuit authors of the seventeenth century. Especially amongst the Protestants, the Exercises soon obtained a dubious reputation – they were suspected to be a special medium that could make people return to the Catholic Church.
244
Angela Mayer-Deutsch
had in Jesuit education and by the large number of books of emblems and imprese. Many Jesuits used emblems and pictures in their missionary work. They believed that a skillfully constructed emblem would mediate religious truth faster and more effectively than a sermon and therefore used it in their struggle for the souls of the pagans. The Jesuits achieved a remarkable propaganda effect with the help of emblems that were inserted into architecture and ceremonial architecture, stage sets, paintings, books of prayer, collections of sermons, religious cards, and demonstrations in museums and palaces. When the centennial of the order in 1640 was celebrated with a book, it was decided that the genre of the publication should not be a primarily historical or panegyrical work, but a book of emblems, Imago Primi Saeculi Societatis Jesu. The book opened with the emblem of the Society: the picture shows the sun shining onto the world and the epigraph says: “Nobody can hide from its warmth,”19 which clearly expresses the Jesuits’ universal claim to power that was equally apparent in their natural philosophy. Early modern experimental natural philosophy was applied by Kircher in accordance with his ‘emblematic’ worldview to comprehend the ‘old’ as well as the ‘new’ world and to rediscover God as a result of this insight. Using a term coined by Catherine Chevalley and transferring it from the history of ideas to the history of collections, one might say that Kircher transformed natural philosophy into a site of ‘permanent conversion.’20 This means that for Kircher every recognition and every explanation was made in the service of God and the miraculous universal divine truth, implying that the explorer was, as it were, ‘permanently converted.’ The Jesuits chose the subjects that were supposed to stimulate their imagination from a wide Christian and mythological context as well as from the imperial iconography of their patrons. They used written images (saying “Gloria Dei, Amen”), the cross, the chalice, the figure of the pilgrim or the shepherd, the figure of Death and of the sinner in pur19 20
“Non est qui se abscondat a calore eius,” quoted in Mario Praz. Studies in Seventeenth-Century Imagery. Vol. 1. London: Warburg Institute, 1939-1947. 170. Catherine Chevalley. “L’ars magna lucis et umbrae d’Athanase Kircher. Néoplatonisme, hermétisme et ‘nouvelle philosophie.’” Baroque 12 (1989): 109. With regard to this convergence of natural philosophy and divine wisdom Chevalley claims that there is a gradual movement of the mind: “Le culte rendu à la sagesse divine consiste alors dans ce mouvement de l’esprit qui commence avec l’étonnement devant la richesse et le mystère des phénomènes, et se poursuit avec le dévoilement de leur structure cachée, géométrique.” Here she is making a point that I shall pick up on in the following passages.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
245
gatory, as well as saints, angels, demons, and subjects from the history of salvation; furthermore they used the double eagle, doves, portraits of popes and emperors, as well as their names as pictorial subjects for optical, magnetic, acoustic, pneumatic, and hydraulic machines. In my opinion, one of the primary practices of the Exercises, namely the imitation of the Vita Christi, was paralleled especially in those subjects that were chosen for the visualization of optical machines: the life, Passion, death, and resurrection of Christ. They appeared either in the form of entire scenes or as single symbols in the pictorial representations applied to sundials, to the Laterna Magica and the Camera Obscura, or to the smicroscopium parastaticum, and also in the “giant crystal globe filled with water that showed the resurrection of Christ within the water.”21 The constitutiones, the rule of the Jesuit order (1540), began as follows: “1. Whoever in our Society, which we wish to honor with the name of Jesus, wants to fight for God under the banner of the cross and to serve the only Lord as well as the Roman pope, his governor on Earth . . .”22 ‘To fight for God under the banner of the cross’ – is what several of the engravings seem to demand, in my opinion. The cross, integrated into the Jesuit coat of arms,23 was one of the most simple as well as one of the most popular motifs in the iconography of optical machines. It was used for the camera obscura as well as for several forms of mirror projections, anamorphoses, and sundials. I would argue that the various engravings on portable sundials were not of much practical use, but achieved their significance by means of a distinct playfulness and illusionism as well as a religious symbolism. The iconographical variety of the Gnomonica Kircheriana ranged from the cross, the solar disc, and the coat of arms of the Society to Death, mathematical forms, or the Habsburg double eagle, and, astonishingly, eggs. Those images primarily represented experiments of thought. Here, we recognize the fictionalism and probabilism of ‘Jesuit natural philosophy.’ We see fictitious, hypothetical forms that have never been realized, and we feel the pleasure of playing with them. Natural philosophy was allowed to play with probabilities; divine truth remained always and necessarily untouched by it.
21 22 23
“Globus crystallinus magnus aqua plenus Salvatoris resurrectionem in mediis aquis repraesentans.” De Sepi. Musaeum. 3. Heinz-Joachim Fischer. Der heilige Kampf. Geschichte und Gegenwart der Jesuiten. Munich: Piper, 1987. 65. The coat of arms of the Jesuits shows a heart pierced by three arrows together with a cross and a monogram of Christ.
246
Angela Mayer-Deutsch
Even before Kircher, the senses, especially the eyes, were supposed to be directly connected with the human heart.24 This spiritual and philosophical path of recognition that leads through the senses to the heart will be traveled in the following passages by looking at selected engravings from Kircher’s Ars Magna Lucis et Umbrae. Different modes of perception will be passed on the way, starting with confusion and sheer wonder continuing via “active contemplation” and recognizing wonder and ending with divine illumination. Confusion and Intimidation In the use of the Laterna Magica playing with fictitious forms was not as important as specific strategies of scaring the audience, which were in many ways similar to certain features of the first week of the Exercises. More than two hundred years before Constantijn Huygens, Thomas Rasmussen Walgenstein, Francesco Eschinardi, or Athanasius Kircher, a book by Giovanni da Fontana, the Bellicorum Liber Instrumentorum (around 1420),25 gives evidence for the three components of the Laterna Magica: a source of light, a picture, and a projection surface. There are only two elements missing, namely a planoconvex lens, the “condensor lens,” as it was called, or the concave mirror to focus the rays, and a further lens to focus the picture. Fontana’s text, which is printed directly below the projected image of a devil, characterizes the Laterna as “a nocturnal appearance to intimidate the observer.” The tradition of experimenting with projections of images originated in ancient times and this textual and visual source, which is technically as well as iconographically interesting, was one of the earliest pieces of evidence for those practices. It is a particularly interesting feature of this text that the choice of the motif of the image is supposed to have the explicit purpose of frightening the observer. In the German edition of his Magia Naturalis (1657), Caspar Schott, a student of Kircher, specifically emphasizes the idea of ‘well-aimed scaring’ in order to discipline oneself: But this presentation of pictures and shadows in dark chambers is much more terrifying than those generated by the sun. With the help of this art godless 24 25
One could name here Thomas Aquinas or Robert Bellarmin. Bayerische Staatsbibliothek Munich Cod. Icon 242, printed in David Robinson. The Lantern Image. Iconography of the Magic Lantern 1420-1880. East Sussex: Magic Lantern Society, 1993. 6, 13.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
247
people could be easily kept from vice if the image of the devil was drawn on the mirror and then projected into a dark place.26
On the one hand, the Christian moral appeal as well as the pleasure of pulling the wool over the naive observer’s eyes (‘Hinters Licht-Führen’), as presented again and again in Kircher’s work, was often combined with a warning against pagan and black magic practices. On the other hand, it was connected with reproaching profit-seeking traveling entertainers, like the one presented in this description of a camera obscura with living actors. Such traveling entertainers attracted curious people who were quite eager to discover secret things in a dark chamber with no light at all shining inside but only as much as a tiny sheet of glass might let pass. Then they sharply advise one to be as silent as a mouse . . . as if one was waiting for the sermon in church. While waiting for the apparition they say that the devil will be here soon. In the meantime one of the assistants puts on the mask of a devil . . . and walks around as if in deep thought towards a place from where his color and figure can shine through the sheet of glass into the chamber . . . After that someone brings in a large board made of paper [Papyrne Tafel] and positions it in the ray of light that has been directed into the chamber. Then there appears an image [Bildnuss] in the shape of the devil walking up and down, which they [the audience] look at, inspect and view with fear. Thus the poor and inexperienced people waste their time and money viewing the shadow of a traveling entertainer.27
Kircher chose pictorial subjects for his optical machines that were equally intimidating and admonishing (fig. 2). For an arrangement of three camerae obscurae in an obelisk he chose a reprint of Christoph Scheiner, which showed the devil represented with a trident, the cross, and the sundial as motifs. The inner wall served as a screen. Accordingly he and Scheiner did not choose motifs from the external world (such as a landscape, a tower, or a tree) as other scholars usually did, but instead they decided to use painted or drawn pictorial motifs for the projections inside the obelisk. And the picture carrier – possibly on purpose – remained unclear in this case. This woodcut, which probably derives from a nigromantic experiment for the entertainment of Kaiser Rudolf II,28 is a reprint of Christoph Scheiner’s Oculus, hoc est Fundamentum Opticum (1619).29 It shows that the theme of fear and admoni26 27 28 29
Caspar Schott. Magia Optica. Das ist Geheime und doch naturmässige Gesichtsund Augenlehr. Würzburg, 1671. 407. Schott. Magia Optica. 181. Cf. Ignacio Gomez de Liano. Athanasius Kircher (1602-1680), Itinerario del extasis, o las imagines de un saber universal. Madrid: Ediciones Siruela, 2001. 339. This “eclectic” practice of citing and reprinting texts and pictures without giving references is not unusual for Kircher. Schott’s letter of 1652 started out with an
248
Angela Mayer-Deutsch
Fig. 2: Three Camerae Obscurae in a model of an obelisk. Athanasius Kircher. Ars Magna Lucis et Umbrae (Rome, 1646 and Amsterdam, 1671).
tion was not a novelty at Kircher’s time and did not end with his work either. Variations on this theme continued into the eighteenth century30 and later. Nonetheless the quantity of such subjects and their use in the
30
allusion to Kircher’s copying word for word from Francois Aguilon’s Opticorum libri sex in a letter dedicated to Cardinal Francesco Barberini for Kircher’s Prodromus Coptus (1636). In order to avoid accusing Kircher, Schott gallantly blames the master’s inattentive assistant for the plagiarism. Poliander’s Anmuthige und Zeit-kürtzende aus vielen curieusen und nützlichen Wissenschaften hergeleitete Quellen. Erster - Fünfzehnter Gang. Erfurt, 1720. 260 shows a tower instead of the obelisk and the figure of the devil, drawn as an expressive caricature of the common figure in shadow theater. The other motifs appear in their usual form. Cf. Bodo von Dewitz and Werner Nekes. Ich sehe was, was Du nicht siehst! Sehmaschinen und Bilderwelten. Die Sammlung Werner Nekes. Göttingen: Steidl, 2002. 51.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
249
Musaeum Kircherianum gives rise to certain questions. The connection between the power of images and fear is quite obvious here and was consciously utilized. Depictions of Hell naturally also belonged to the repertory of images with the power of causing fear, since Hell, with its origins in the Holy Bible, was a reality for the Christians of that time. It is the place where not even the loving God can reach the closed heart of man any more. At the end of a long tradition of infernal poetry and depictions of Hell, such images finally became part of satires and were also used in the Laterna magica like, for example with Kircher, the well-known picture of a soul languishing in purgatory. The first week of the spiritual exercises was meant to frighten the exercitant and lead him to a mental state of self-humiliation. The guideline for the “contemplation of Hell” was supposed to effect all senses. In a first prelude “the length, width, and depth” of Hell should be contemplated within the imagination. In a second prelude “an inner feeling of punishment . . . which is experienced by the damned” should be suffered. In accordance with the hierarchy of the five senses the exercitant should then exhaust the sensual potential of Hell in five stages, using his imagination.31 The most important aspect seems to be the intimidating effect of certain images, which was imposed upon the excercitant as an “inner feeling of the punishment” that was endured by those hopelessly damned figures shown in agony, screaming, with outstretched arms. This effect was above all achieved by an appeal to the visual and the olfactory sense. The latter was stimulated by the smoke of a dim oil lamp in the room, which was possibly one of the major reasons why this particular subject was chosen from a series of eight pictures on the slide. Kircher actually used the term Lucerna Magica which refers to the underworld or, more precisely, to the ‘lucerns’ or little oil lamps in the catacombs. Suffering the infernal pains leads through death towards the subject of the resurrection. The second engraving in Ars Magna Lucis which illustrates the use of the Laterna Magica, shows the figure of 31
“The first point should be: to behold the giant embers and the souls that seem to be caught in burning bodies by the power of imagination. The second point should be: to hear with one’s ears the sounds of moaning, howling, the screams and blasphemous words against Christ our Lord and all his saints. The third: to smell with the sense of smell smoke, sulfur, filth, and rotting things. The fourth: to taste with the sense of taste bitter things like tears, sadness, and the worm of conscience. The fifth: to touch with the sense of touch in the way the embers touch the souls and burn them.” Peter Knauer. Ignatius von Loyola. Geistliche Übungen und erläuternde Texte. Graz: Styria, 1978. 44.
250
Angela Mayer-Deutsch
Death projected on a wall. Thus, the primary exercise of the first week (consternation, intimidation, and suffering) was depicted several times in the iconography of optical machines. Wonder This specific use of pictures in the museum was probably influenced by contemporary mystery plays and the propagandistic Jesuit theater which used drastic and powerful devices like crowd scenes, choirs, etc. to stimulate the imagination of the audience. The stage construction included methods of capturing ghostly apparitions and machines to produce clouds, thunder, and wind. The pictures that were shown in such performances achieved their particular significance as a means of instruction and excitement for the visitors and aimed at strengthening their faith or converting them by appealing to their senses. In 1625, the young Kircher also contributed to stage productions and pyrotechnical demonstrations that were performed at the college in Heiligenstadt in honor of the Prince Elector of Mainz, in the service of whom he remained for a short while. In his autobiography he says: I was entrusted with the presentation of scenic acts. The legates watching were struck with amazement by the extraordinary things that happened on stage. There were also people who accused me of the crime of black magic (Zauberei) and others who cast even worse aspersions upon my character. Thus, to remove the suspicion of that crime, I was forced to explain my whole procedure in detail to those legates. To their great content my explanation was so entirely convincing that they grew very fond of me.32
In their stupor, the spectators demanded an explanation. The amazement could be transformed into a knowing amazement or wonder.33 From the Kircherian perspective, this was achieved by “active contemplation” and a more or less detailed knowledge of the basic mechanism. The second week of the Exercises, which followed the stages of consternation, the sharpening of the senses, and self-humiliation, addressed the contemplation of the incarnation of God within the human soul. According to Kircher, God’s descent to the realm of man was perceptible 32 33
Seng. Selbstbiographie. 25. See the various publications of Lorraine Daston for a detailed history of wonder, attention, and curiosity. I only mention here Lorraine Daston and Katharine Park. Wonders and the Order of Nature, 1150-1750. New York: Zone Books, 1998 and Lorraine Daston. Eine kurze Geschichte der wissenschaftlichen Aufmerksamkeit. Munich: Carl Friedrich von Siemens Stiftung, 2001.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
251
in every earthly phenomenon, since those were believed basically to imitate divine creation. Man’s duty, therefore, was the “active contemplation” of this mirror. This Ignatian maxim was exemplified in Kircher’s entire work. Jeronimo Nadal, S.J., Ignatius’s early companion, described it as follows: “He contemplated in all things, actions, and conversations the presence of God and experienced the reality of spiritual things, so that he was a contemplative in action, so to speak (a theme he commented on as follows: God has to be found in everything).”34 The path leading to God went from his incarnation in Jesus through the latter’s life, Passion, death, and, finally, his resurrection. The exercise called “contemplation of two banners” addressed the sensual representation of the empire of Satan and the kingdom of Christ. According to Ignatius, man was the arena where Christ and Satan competed for man’s participation in their work. By contemplating the “two banners,” the exercitant should arrive at a certain sensitivity to or intuition of the tricks and deceits of Satan on the one hand and Christ’s intentions on the other hand. As Kircher explains in Oedipus Aegyptiacus, this exercise is again comparable to the attempt to delude superstitious people with demonstrations of his machines.35 The aim of this was to show them quite plainly the dubious nature of pagan priests and their practices. As an example Kircher mentions the speaking statues of pagan priests, which were used to force ignorant people to oblation by means of the trick of a speaking tube. Despite this he himself devoted an entire chapter of the museum’s catalogue, entitled “Delphic oracle,” to the speaking tube, and he enjoyed showing the illusion to visitors to the museum. Ultimately, the pictorial demonstration of the non-functioning of certain machines, especially the Perpetua Mobile machines, emphasized the limits of the analogy between a human Magus and the omnipotent Creator. Real miracles resisted any kind of demonstration and replication, and remained unaffected by such playful nonsense. Again we notice the probabilistic element of ‘Jesuit natural philosophy,’ the pleasure of playing with hypotheses, the truth of which was not relevant!
34 35
Quoted in Joseph F. Conwell. Contemplation in Action. A Study in Ignatian Prayer. Spokane: Gonzaga University Press, 1957. 25. Athanasius Kircher. Oedipus Aegyptiacus . . . 2 vols. Rome, 1653. Vol. 2, 337f.
252
Angela Mayer-Deutsch
Transformation The Theatrum catoptricum octogonale focused on the perceptive and psychic metamorphosis of the spectator effected by the refraction and reflection of light rays (fig.3). This confused the spectator, who saw his own image and pictures (of a sun, a donkey, a lion, etc.) alternating in the mirror, an effect that was achieved by the fast rotation of an octagonal drum of pictures. The result of the performance was probably not only a transformation of his perception but of his psyche as well. Kircher writes: I own such a machine myself, and it caused great amazement in everyone who beheld the face of a wolf, a dog, or some other animal instead of his own natural face.36
This astonishing metamorphosis, as I would argue, was not only supposed to perplex and entertain the excercitant, it was also intended to initiate the contemplation of the enigma of incarnation and reflection on his own forthcoming transformation. The theme of the third week of the Exercises was the mystical tradition, the life and death of Christ, which again occurs as a motif in the Ars Magna Lucis when the picture of a cross is used in the Laterna Magica and in the smicroscopium parastaticum (fig.4), as it was called – a predecessor of the phenakistiscope. A slide bearing the pictures of the Passion cycle was placed between the two plates of the ‘microscope.’ Using a tube, which contained a lens and which at the same time served as a crank handle for the upper plate, one image after another was contemplated through a hole in the upper plate. Thus the mechanism involved an element of movement, too. Even though this aspect of movement was again perceived by the senses, it aimed at the same time at the movement and expansion of the psyche. Illumination Finally, the resurrection of Christ and the process of the illumination of the exercitant, which were the themes of the fourth and final week, also became part of the museum’s catalogue and other works of the Jesuit. For Kircher, the Biblical resurrection was ‘evidenced’ by means of analogy by the “palingenetic” experiment already described, furthermore it 36
Athanasius Kircher. Ars Magna Lucis et Umbrae . . . Amsterdam, 1671. 901.
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
253
Fig. 3: Cylindric anamorphosis and “Theatrum Catoptricum Octogonale” from Athanasius Kircher. Ars Magna Lucis et Umbrae (Rome, 1646 and Amsterdam, 1671).
was depicted in the highlight of the museum that was also mentioned above, namely in the “giant crystal globe filled with water which showed the resurrection of Christ within the water.”37 In the opinion of a devoted Catholic like Kircher, this engraving represented the divine light (lux) as it shone through the window of a room 37
De Sepi. Musaeum. 3.
254
Angela Mayer-Deutsch
Fig. 4: “Smicrocopium Parastaticum” from Athanasius Kircher. Ars Magna Lucis et Umbrae (Rome, 1646 and Amsterdam, 1671).
Fig. 5: Prismatic reflections from Athanasius Kircher. Ars Magna Lucis et Umbrae (Rome, 1646 and Amsterdam, 1671).
Musaeum Kircherianum and the Ignatian Exercitia spiritualia
255
and was then refracted (umbra, which means ‘refracted light’) by two prismatic mirrors that are shown in detail in the upper part of the copperplate engraving. As a result, the right corner of the cubiculum is illuminated by the manifold forms and colors that result from this refraction, among them faces or masks and stars (fig. 5). The refracted light provided man with the vision of the divine light. This can be interpreted as a pictorial realization of the visionary, the illumination that was part of the Exercises. In that way the cubiculum would symbolize the heart of the exercitant, enflamed with love of God. I would argue that, using “active contemplation,” the ideal visitor to the museum would embark on a spiritual process of perception and cognition. After a state of frightened confusion, ignorant amazement, and wonder he would advance to a kind of informed admiration of divine creation, as presented in various demonstrations in the ideal museum, and he would acquire an understanding of its basic mechanism. Finally he would achieve the highest ‘illuminated’ state, perceivable only in a momentary and visionary manner, the recognition of divine truth. This is where the ways of the heart end. Translation: Kathrin Bethke
WORKS CITED Ashworth, William B., Jr. “Natural History and the Emblematic World View.” Reappraisals of the Scientific Revolution. David C. Lindberg and Robert S. Westman. Cambridge: Cambridge University Press, 1990. 303-32. Boyle, Robert. The Works. Ed. Thomas Birch. Hildesheim: Olms, 1966. Chevalley, Catherine. “L’ars magna lucis et umbrae d’Athanase Kircher. Néoplatonisme, hermétisme et ‘nouvelle philosophie.’” Baroque 12 (1989): 95-109. Conwell, Joseph F. Contemplation in Action. A Study in Ignatian Prayer. Spokane: Gonzaga University Press, 1957. Daston, Lorraine and Katharine Park. Wonders and the Order of Nature, 1150-1750. New York: Zone Books, 1998. Daston, Lorraine. Eine kurze Geschichte der wissenschaftlichen Aufmerksamkeit. Munich: Carl Friedrich von Siemens Stiftung, 2001. Dewitz, Bodo von and Werner Nekes. Ich sehe was, was Du nicht siehst! Sehmaschinen und Bilderwelten. Die Sammlung Werner Nekes. Göttingen: Steidl, 2002. De Sepi, Giorgio. Musaeum celeberrimum. Amsterdam, 1678. Fischer, Heinz-Joachim. Der heilige Kampf. Geschichte und Gegenwart der Jesuiten. Munich: Piper, 1987. Gomez de Liano, Ignacio. Athanasius Kircher (1602-1680), Itinerario del extasis, o las imagines de un saber universal. Madrid: Ediciones Siruela, 2001.
256
Angela Mayer-Deutsch
Georges, Heinrich. Lateinisch-Deutsches Handwörterbuch. Hanover, 1892. Jacobus de Voragine. Legenda aurea. Jena: Diedrichs, 1925. Kircher, Athanasius. Ars Magna Lucis et Umbrae . . . Rome, 1646 and Amsterdam, 1671. Kircher, Athanasius. Oedipus Aegyptiacus . . . 2 vols. Rome, 1653. Kircher, Athanasius. Arca Noê . . . Amsterdam, 1675. Knauer, Peter. Ignatius von Loyola. Geistliche Übungen und erläuternde Texte. Graz: Styria, 1978. Poliander. Anmuthige und Zeit-kürtzende aus vielen curieusen und nützlichen Wissenschaften hergeleitete Quellen. Erster - Fünfzehnter Gang. Erfurt, 1720. Praz, Mario. Studies in Seventeenth-Century Imagery. Vol. 1. London: Warburg Institute, 1939-1947. Robinson, David. The Lantern Image. Iconography of the Magic Lantern 1420-1880. East Sussex: Magic Lantern Society, 1993. Schott, Caspar. Magia Optica. Das ist Geheime und doch naturmässige Gesichts- und Augenlehr. Würzburg, 1671. Seng, Nikolaus. Selbstbiographie des P. Athanasius Kircher aus der Gesellschaft Jesu. Fulda: Druck und Verlag der Fuldaer Actiendruckerei, 1901. Villalpando, Juan Battista. Explanationes in Ezechielis et apparatus urbis ac templi Hierosolymitani. Rome, 1604.
CONNY RESTLE
Organology: The Study of Musical Instruments in the 17th Century Musical instruments have been among the most important artifacts since Greek antiquity. The oldest representations of musical instruments and scenes with musical content come from the early Cycladic period (2700-2100 B.C.). It appears that there was no word for music then or in Homer’s time. Only at the beginning of the fifth century B.C. does Pindar name music in his first Olympic ode. At this same point, music schools were founded in Athens, devoted to the education of usually well-off youths in music theory, singing, and the playing of instruments.1 The word “organon,” meaning tool or instrument, was first used in a musical sense by Vitruvius in the third century B.C. in his book De architectura.2 He describes how the Greek engineer Ktesibios of Alexandria built a particularly complicated instrument with numerous pipes, which could be provided with air through an inventive interplay of water and various pumps and chambers. Although Vitruvius calls this instrument “hydraulis,” the more comprehensive term “organon” then became the standard, and in the course of late antiquity the term came into use as pars pro toto for all kinds of musical instruments. The history of the term “organon” is actually worth investigation in itself, and for that reason will not be foregrounded here. The topic here is rather what was understood in the seventeenth century as a “musical instrument,” that is “organon,” and with what means theorists approached the phenomenon of works of art that incorporate elements of sound within them. A brief word about the historical position of music around 1600: Since the middle of the sixteenth century, instrumental music had been 1
2
On this topic, cf. Ministerium für Kultur der Republik Griechenland, ed. Geschenke der Musen. Musik und Tanz im antiken Griechenland. Athens: Ministerium für Kultur der Republik Griechenland, 2003. Cf. Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard University Press. 1962. Vol. 2, 311-19.
258
Conny Restle
raised to the status of an artistic genre, along with vocal music. Wealthy ruling houses in Italy, France, and Germany maintained their own instrumental ensembles (orchestras) and to this end laid the groundwork for significant collections of musical instruments in their art cabinets.3 This led in turn to the foundation of numerous instrument construction workshops, particularly in Italy (Venice, Rome, and Florence). North of the Alps, traditional guild regulations initially prohibited the inclusion of instrument makers, until it had to be acknowledged that there was a flourishing trade in the construction of instruments. Around 1600 most instruments were arranged in groups, so-called families, similar to the registers of a vocal ensemble, consisting of bass, tenor, alto, and soprano or discant. With the advent of opera and the increasing importance of solo singing, individual instruments were also refined in sound and playing technique. Researchers and theorists – from the field of music as well as the natural sciences – began to describe and analyze music systematically, because of the great interest from musicians and composers in special compendia.4 Galileo Galilei, for example, who came from a highly musical family in Pisa – his father Vincenzo was a famous lute player and assisted in his son’s acoustic experiments (Galileo himself played the lute very well and enjoyed composing for his instrument) – conducted attempts around 1600 to determine intervals and their rate of vibration by means of strings and weights. In doing so, he was continuing experiments from antiquity that he had learned of during his studies at the University of Pisa. Additionally, he adapted the arrangement of the frets on the neck of a lute in order to measure the acceleration of bodies on inclined planes. Neither Johannes Kepler nor Galileo worked on a comprehensive description of physical and aural qualities, because they were interested in fundamental research rather than in the very specific craft and technique of musical instruments. It was Michael Praetorius, one of the most outstanding and wellknown composers of his time (he was chamber organist of the Episcopal Halberstadt court orchestra in Groningen, court orchestra director of the ducal orchestra of Braunschweig-Lüneburg, as well as chamber master of the Magdeburg archiepiscopal court orchestra at Halle an der Saale and the Saxon electoral orchestra at Dresden), who was the first to di3 4
Julius von Schlosser. Die Kunst- und Wunderkammern der Spätrenaissance. Ein Beitrag zur Geschichte des Sammelwesens. Leipzig: Klinkhardt und Biermann, 1908. On the study of instruments in general, cf. John Henry van der Meer. “Instrumentenkunde.” Die Musik in Geschichte und Gegenwart. Ed. Friedrich Blume et al. 2nd rev. ed. Kassel: Bärenreiter, 1996. Vol. 4, 951-70.
Organology
259
Fig. 1: Frontispiece of Michael Praetorius. Theatrum Instrumentorum (Wolfenbüttel, 1620 [1619]).
rect his work specifically at musicians and instrument makers. He dedicated his De Organographia,5 which he wrote in German, “To all organists, instrumentalists, organ and instrument makers and those, not only Teutons but also those of other nations, who exercise and admire the Musicam instrumentalem” (fig. 1). The title page reads as follows: Including all musical instruments old and new, foreign, barbarian, primitive, and obscure, as well as native, artistic, pleasant, and familiar. Their nomenclature, intonation, and characteristics, their correct representations and precise reproductions, useful, necessary, and essential not only for organists, instrumentalists, organ and instrument makers, but also entertaining and delightful to read for philosophers, philologists, and historians. Contains a complete register. 5
Michael Praetorius. Syntagma Musicum. De Organographia. Wolfenbüttel, 1619 and 1620.
260
Conny Restle
It is true that even before Praetorius, at the beginning of the sixteenth century, there was a smaller book about musical instruments by Sebastian Virdung entitled Musica getutscht.6 This book, however, contained only a listing of the commonly used instruments, partially illustrated with woodcuts. Praetorius cites Virdung, though he does say that there is nothing to be found in his book about the use and the qualities of the instruments, above all their sonorific and technical qualities: A book printed in quarto in Basel in 1511 anno Christi can be found in quite a few libraries, in which quite a few of the old and well as many of the current instruments are depicted.7
It must be said at this point that as of yet no basis for Praetorius’s work has been discovered, so that Praetorius can be taken as the first to systematize modern music history. He himself decided on the following arrangement: First, the wind instruments, beginning with the organ, then the brass and woodwinds. These are followed by the percussion instruments. Finally Praetorius turns his attention to the stringed instruments, both bowed and plucked (fig. 2). What idea governs this systematization? It is their method of producing sound. “As for the classification of such musical instruments/one cannot justifiably separate them from each other/except by their tone and sound.”8 At the same time, playing techniques are also taken into account, the instruments being organized by family and pitch. Not until Hector Berlioz’s instrumentation manual, which the composer and music writer published in Paris in 1843/44 and expanded in 1855,9 was there any comparable treatise on musical instruments. This may be surprising, but it shows that over the course of almost 250 years, virtually nothing essential changed in the instrumentation of the orchestra. It is Berlioz then, a good seventy years after the rise of the classical orchestra in Mannheim, who is the first to analyze and systematize the changed situation. But back to Praetorius. It is astonishing that Praetorius takes the term “instrument(s)” for granted. He uses it without reflecting on whether other “tools” (e.g. medical or technical) could possibly be understood by this term. He even stresses that his knowledge is the result of his “own diligent research and experience”10 and, remaining true to his po6 7 8 9 10
Sebastian Virdung. Musica getutscht. Basel, 1511. Praetorius. Syntagma Musicum. 5v. Ibid. 1. Hector Berlioz. Grand traité d’instrumentation et d’orchestration modernes [1843/ 1844]. Paris: Schonenberger, 1855. Praetorius. Syntagma Musicum. 5r.
Organology
261
Fig. 2: Systematization of musical instruments from Michael Praetorius. Theatrum Instrumentorum (Wolfenbüttel, 1620 [1619]).
sition as a working musician, that there are no special secrets in the area of musical instruments, “or in Nature’s hidden reasons and causes.” For him, no questions of physics arise as to the form of sound and pitches. It is quite different with Marin Mersenne. Next to Thomas Hobbes, René Descartes, Athanasius Kircher, Robert Fludd, and Johannes Kepler, he was one of the great polymaths of his time. He was educated at the Jesuit College of La Flêche in logic, physics, metaphysics, mathematics, and theology. He continued his theological studies at the Sorbonne. He entered a monastery in 1611 and was ordained as a priest in 1612. Beginning in 1619 he taught in Paris, where he founded a kind of academy of sciences in 1640.
262
Conny Restle
His most important work is the Traité de l’Harmonie universelle. The first volume of the Latin version Harmonicorum libri appeared in French in 1627.11 In this text, Mersenne represents what Walter Wiora once called the “theory of music as a branch of autonomous knowledge.” Everything is directed toward recognizing the existence of God in his works and creations. By harmony, Mersenne understands everything that can be recognized as a regulated order. And in all of arithmetic, there is nothing so useful as the numerical ratios found in music. Many of Mersenne’s findings were initially attributed to other physicists, such as the observation that the range of a soundwave increases according to its direction or his thoughts on overtones. He mentions the correct explanation of overtones in strings, but then he dismisses it. He does, however, correctly explain overtones in air columns. In doing so, he was the first to define sound and pitch as vibration. In contrast to Descartes, he rightly recognized that high and low, strong and weak pitches traveled equally quickly. Mersenne’s thoughts on tempering, that is, on the tuning of musical instruments, were especially important in the following decades. He calls for a twelve-step, evenly tempered system, since deviation from the pure intervals is so small that the human ear does not experience it as noticeable or disturbing. He evens mentions the possibility of a tempered quarter tone system (first realized in the twentieth century by Alois Hába). The section covering the musical instruments known and used in Mersenne’s time spans seven of the total of seventeen books of the Harmonie universelle and as such takes up the largest amount of space in the book.12 Marini Mersenni . . . Harmonicorum libiri, in quibus agitur des sonorum natura, causis, et effectibus: de Consonantiis, Dissonantiis, Rationibus, Generibus, Modis, Cantibus, Compositione, orbisque totius Harmonicis Instrumentis. Opus utile Grammaticis, Oratoribus, Philosophis, Iurisconsultis, Medicis, Ma-thematicis, atque Theologis.
Like Praetorius, Mersenne describes all the instruments in common use. His descriptive technique is different than that of Praetorius, however, because he always takes an instrument pars pro toto out of its family and then reviews it thoroughly. Once again, the instruments are ar11 12
Marin Mersenne. Harmonie universelle. Ed. François Lesure. Paris: Centre national de la recherche scientifique, 1963 [facsimile of the edition Paris, 1636]. Marin Mersenne. “Traité des instrument à cordes.” Harmonie universelle. Cf. the English edition, Harmonie universelle. The Book on Instruments. Trans. Roger E. Chapman. The Hague: Nijhoff, 1957.
Organology
263
ranged according to the way they are played, but Mersenne subdivides the stringed instruments very precisely, while there is no similarly meticulous classification for the woodwinds and percussion instruments. And in contrast to Praetorius, Mersenne also explains problems of music theory, like, for example, the proportions of intervals and their realization on a monochord. In one important point Mersenne thoroughly supersedes Praetorius: the relationship of nature and art in the construction of musical instruments. Mersenne comes to the conclusion that instruments were indeed created after the model of nature, but that later they were developed further according to their own rules. This is the question: “Determiner si l’on a fait les Instrument de Musique à l’imitation des voix, ou si l’on a reglé les intervalles des voix par ceux des Instruments.”13 To determine if musical instruments were made in imitation of the voice, or if the intervals of the voice were determined by those of the instruments. And consequently, to determine if art perfects nature, or nature art, and if we can compare the creations of art with the creations of nature. Furthermore, the examples of notation that Mersenne adds to the descriptions of the instruments are interesting. It is clear that in contrast to Praetorius, he wrote his treatise not only for musicians and music scholars, but also for a wider audience of scholars as well as the interested general reader. Mersenne’s Harmonie universelle is in sum the first treatise that represents the musical instrument not just as a mere tool for the creation of music, but which understands the musical instrument as an autonomous technical object of art, which can be represented by mathematical and physical formulas. The third great intellectual figure in early studies of musical instruments is Athanasius Kircher, born in Geisa in the Rhön. He was educated by his father, who was professor of theology and magistrate at Fulda. At seventeen he had already become a novice and had joined the Jesuits. Alongside theology and the usual subjects, he was interested in logic, physics, and philosophy, as well as oriental studies and oriental languages. In 1629, he was offered a professorship in mathematics, philosophy, and oriental languages at Würzburg. A few years later he went to Avignon and was named court mathematician to Emporer Ferdinand II. While he was traveling to Vienna via Rome, he was offered a position at the Collegium Romanum. This was a stroke of luck for Kircher, because he could work there in the very center of the current intellectual world and with the most significant scholars. His yearning for a uni13
Ibid. 7, III. Proposition.
264
Conny Restle
Fig. 3: Anatomy of the human voice from Athanasius Kircher. Musurgia universalis (Rome, 1650).
versal discussion of scholarly problems, however, frequently resulted in an unsystematic treatment of the material. In his Musurgia Universalis sive Ars Magna Consoni et Dissoni in X. libros digesta,14 published in Rome in 1650 (and which appeared in a partial German translation by Andreas Hirsch of Bächlingen in 1662 in Schwäbisch Hall), Kircher attempts an all-encompassing, almost cosmological presentation of music and its tools, that is, of musical instruments. He goes to great efforts to bring together the totality of the musical knowledge of his time. In order to do so, Kircher begins with the physical processes that create pitch, including the anatomy of human hearing and the voice (fig. 3). Following this, Kircher gives a historical overview of music and instruments and describes them in terms of mathematics and physics (fig. 4). Turning to the language of music and 14
Athanasius Kircher. Musurgia universalis Musurgia universalis sive ars magna consoni et dissoni in X. libros digesta. Hildesheim and New York: Olms, 1970 [facsimile of the edition Rome, 1650].
Organology
265
Fig. 4: How soundwaves travel from Athanasius Kircher. Musurgia universalis (Rome, 1650).
266
Conny Restle
Fig. 5: Fantastic organ from Athanasius Kircher. Musurgia universalis (Rome, 1650).
the teaching of composition, Kircher then comes to the teaching of affect, widely discussed in his time. In between, in the sixth book, “liber organicus,” he thoroughly discusses the problems of the study of instruments. He continually concentrates on the representation of acoustic problems. The ninth book, “liber magicus,” with its fantastic descriptions of the miracles and secrets of the art of pitch, is particularly interesting on this subject. He concludes, in the tenth book, with the harmony of the spheres. The wealth of material with which Kircher confronts his readers culminates in a final, magnificent, baroque representation of medieval scholastic terms of scholarship, and was meant to revive the Artes liberales. The whole range of musical instruments is presented in Kircher’s work as a fixed structure that can hardly be perfected or changed in any way. If there are particularities, these are automatic musical instruments or acoustically optimized instruments (fig. 5).15 Kircher, however, does not 15
Ibid. Liber IX, 308ff.
Organology
267
reflect on the radical change that was occurring in instruments at the time, a change that examined the individual form of each instrument and that demanded an intensive and physical (in the modern sense) consideration of the complex acoustic and physical form of the instruments. From Praetorius, through Mersenne and on through to Kircher, scholarly engagement with musical instruments becomes increasingly lively. While Praetorius limited himself entirely to the practice of playing and to general artisanal details, but was the first to introduce a system and to discover the sonority of instruments, Mersenne attempts a mathematical-physical description and explanation of the instruments. Kircher places the discoveries made by Praetorius and Mersenne within the context of Catholic teachings on creation and encyclopedically consolidates all the known discoveries of his time. The study of instruments, organology, had been born as an autonomous discipline. What came after the great music theorists of the seventeenth century? In 1698 Bartolomeo Cristofori in Florence invented hammer mechanics and the pianoforte. And by so doing, he changed the musical world at a stroke. Cristofori’s knowledge was completely different from that of earlier instrument makers. He was an instrument maker as well as restorer, and he held an official position in the courts of the Medici with a regular income and his own budget. Under these conditions, it was possible for him to develop, to discover what was then a completely new instrument.16 Today we know that Cristofori was intensely concerned with geometric and mechanical details and improved these through numerous experiments. Cristofori is primarily known for the invention of the pianoforte, which he had promised to the young prince Ferdinando. In 1688, when he was 33, Cristofori came to Florence, and in 1687 Isaac Newton had published his Principia mathematica, in which he explained the laws of mechanical movement.17 The Medici library contains one of the first copies of Principia mathematica, so that one can justifiably assume that Cristofori was able to study the work in depth. Even Scipione Maffei, the first journalist who reported on Cristofori’s invention,18 knew Newton’s concepts quite well since he was a member of several scholarly academies. Newton’s second law, in 16
17 18
Conny Restle. Bartolomeo Cristoforo und die Erfindung des Hammerclaviers. Munich: Editio Maris, 1990; “Cristofori, Bartolomeo.” Die Musik in Geschichte und Gegenwart. Vol. 5, 97-103. Isaac Newton. Philosophia naturalis principia mathematica. London, 1687. Scipione Maffei. “Nuova invenzione d’un gravecembalo col piano e forte, aggiunte alcune considerazioni sopra gli strumenti musicali.” Giornale de’ letterati d’Italia 5 (Venice, 1711): 144-59.
268
Conny Restle
particular, which describes the acceleration of an object by a force, is applied at several points in Cristofori’s mechanics. Cristofori and his invention represent a completely new phase of musical instrument construction. The instrument is now the result of a modern, intensive achievement in engineering, which the study of instruments will only be able to classify some three hundred years later.19
Translation: Daniel Hendrickson WORKS CITED Berlioz, Hector. Grand traité d’instrumentation et d’orchestration modernes [1843/44]. Paris: Schonenberger, 1855. Kircher, Athanasius. Musurgia universalis sive ars magna consoni et dissoni in X. libros digesta. Hildesheim and New York: Olms, 1970 [facsimile of the edition Rome, 1650]. Maffei, Scipione. “Nuova invenzione d’un gravecembalo col piano e forte; aggiunte alcune considerazioni sopra gli strumenti musicali.” Giornale de’ letterati d’Italia 5 (Venice, 1711): 144-59. Meer, John Henry van der. “Instrumentenkunde.” Die Musik in Geschichte und Gegenwart. Ed. Friedrich Blume et al. 2nd rev. ed. Kassel: Bärenreiter, 1996. Vol. 4, 951-70. Mersenne, Marin. Harmonie universelle. Ed. François Lesure. Paris: Centre national de la recherche scientifique, 1963 [facsimile of the edition Paris, 1636]. Mersenne, Marin. Harmonie universelle. The Books on Instruments. Trans. Roger E. Chapman. The Hague: Nijhoff, 1957. Ministerium für Kultur der Republik Griechenland, ed. Geschenke der Musen. Musik und Tanz im antiken Griechenland. Athens: Ministerium für Kultur der Republik Griechenland, 2003. Newton, Isaac. Philosophia naturalis principia mathematica. London, 1687. Praetorius, Michael. Syntagma Musicum. De Organographia. Wolfenbüttel, 1619 and 1620. Restle, Conny. Bartolomeo Cristofori und die Erfindung des Hammerclaviers. Munich: Editio Maris, 1990. Restle, Conny. “Cristofori, Bartolomeo.” Die Musik in Geschichte und Gegenwart. Ed. Friedrich Blume et al. 2nd rev. ed. Kassel: Editio Maris, 2001. Vol. 5, 97-103. Restle, Conny. “Idee und Realisierung. Bartolomeo Cristoforis Weg zum Hammerflügel.” Jahrbuch des Staatlichen Instituts für Musikforschung (2004): 110-22. Schlosser, Julius von. Die Kunst- und Wunderkammern der Spätrenaissance. Ein Beitrag zur Geschichte des Sammelwesens. Leipzig: Klinkhardt und Biermann, 1908. Virdung, Sebastian. Musica getutscht. Basel, 1511. Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard Univerisity Press, 1962.
19
Conny Restle. “Idee und Realisierung. Bartolomeo Cristoforis Weg zum Hammerflügel.” Jahrbuch des Staatlichen Instituts für Musikforschung (2004): 110-22.
ANDREAS MEYER
In Sound Similar to the Harps: Early Descriptions of African Musical Instruments
When reconstructing the musical cultures of Africa that have been passed on orally, the sources are archaeological evidence and the lyrics of songs that have been handed down, but above all collections and descriptions by European travelers.1 Most travelers, however, had little to do with music; they were missionaries, traders, and later also colonial civil servants and members of the military. The reliability of the descriptions and the accuracy of the documentation vary just as much as their value judgments. Nevertheless, those sources are in many cases historically relevant, especially when combined with the knowledge of present-day music. The particular tone, moreover, tells us something about individual experience, tolerance, and the influence of the Zeitgeist. I would like to show this with the help of some sources from the seventeenth century that concern the music and musical instruments of various African and African American areas. In 1609 the Portuguese priest João dos Santos, missionary of the Order of St. Dominic, published his book Ethiopia Oriental. He had visited various areas in southeast Africa during the years 1586 to 1597. In his book one can find a description of a musical instrument called “Ambira,” with narrow “rods of iron” arranged in a row. The description shows goodwill and even admiration: “And they play so easily, as a good harpsichord player does on the keys; in such a way that the iron segments . . . all together produce a harmony of gentle and soft music.”2 1
2
For example, cf. Walter Hirschberg. “Early Historical Illustrations of West and Central African Music.” African Music 4.3 (1969): 6-18; Gerhard Kubik. Kalimba, Nsansa, Mbira – Lamellophone in Afrika. Berlin: Museum für Völkerkunde, 1998. 75ff.; Christian Läpple. “Ghanaische Musik im Spiegel der Literatur des 19. und 20. Jahrhunderts.” Weltmusik II. Ed. Peter Ausländer and Johannes Fritsch. Cologne: Feedback Studio Verlag, 1983. 3-18. Frei João dos Santos. Ethiopia Oriental [1609]. Lisbon, 1891. 74-75.
270
Andreas Meyer
Fig. 1: Mbira Dzavadzimu. Lamellophone from Zimbabwe.
The instrument belongs to the group of lamellophones, which are only played in African and African American cultures. They consist normally of several iron or wooden lamellas, fixed to a small board or box. The player holds the instrument with both hands and plays the lamellas with his thumbs. In the older literature one therefore sometimes finds it called “thumb piano.” In America they also invented bigger, deep sounding models, where the player sits on the resonating body. It can be assumed that the “Ambira” described by Santos is a forerunner of the Mbira, a lamellophone of the Shona in Zimbabwe. The best known type, however, the Mbira Dzavadzimu (Mbira of the ancestor’s spirits), which is often played in ritual contexts, does not have nine “rods” in a row as described by Santos but two rows with a total of 22 to 24 lamellas (fig. 1). But single row instruments are still known today in southern Africa. In 1620 the second volume of the monumental work of musical theory Syntagma musicum by Michael Praetorius was published.3 This volume deals with musical organology and has an extensive illustrated section. Besides many pictures of European objects, two tables of illustrations have been allocated to instruments from outside Europe (fig. 2 3
Michael Praetorius. Syntagma musicum. Vol. II. De organographia. Wolfenbüttel, 1620.
In Sound Similar to the Harps
271
Fig. 2: Musical instruments documented in Michael Praetorius. Syntagma musicum. Among others a bowed lute (1) and an arched harp (2) from central Africa.
and 3). They are artefacts from contemporary collections. It seems that Praetorius did not have any documentation. Most instruments are called “Amerindian” (“indianisch”) in the captions. In the first table one can find two African string instruments. “Amerindian instruments,” writes Praetorius, “in sound similar to harps.” One of the drawings (fig. 2, no. 2) shows a bowed lute, an instrument only found in sub-Saharan Africa. Each string is stretched over its own tailpiece, which is fixed onto the resonating body. These instruments are predominantly to be found across central Africa. The second illustration shows what is known in organology as an arched harp. The strings are fixed between the upper cover of the resonating body and a curved neck. This type of instrument is familiar in many African regions, but in this specific shape with a round resonat-
272
Andreas Meyer
Fig. 3: Musical instruments documented in Michael Praetorius. Syntagma musicum. Among others a side blown flute (4) from West Africa and an African double bell (5).
ing body and the spaciously carved support for the bow is again only known in central Africa. Praetorius’s second table of illustrations shows among others a side blown trumpet (fig. 3, no. 4) and a double bell (fig. 3, no. 5). The double bell is an instrument commonly found in western and central Africa. “Made from iron and played like our kettle drums,” the caption reads. The side blown trumpet is an engraved elephant tusk (caption: “Amerindian horn of ivory”). Trumpets of this type mainly promote the representation of the status of dignitaries in western and central Africa. The instrument shown is carved at the top. It depicts a crocodile swallowing a man of African origin (wearing a European hat!). It is possible that it was produced especially for export to Europe.4 Similarly decorated in4
Alexander Pilipczuk. Elfenbeinhörner im sakralen Königtum Schwarzafrikas. Bonn: Verlag für Systematische Musikwissenschaft, 1985. 15f.
In Sound Similar to the Harps
273
struments from the West African coast of Guinea are kept in several European ethnological museums.5 Early evidence for a type of instrument common in many regions of sub-Saharan Africa, the xylophone with resonating bodies made of calabash, can be found in the work of Girolamo Merolla, an Italian Capuchin monk who spent time in Luanda in what is now Angola. The description turns out to be as appreciative as the report by Santos on the “Ambira”: One of the favorite instruments is the Marimba. Sixteen calabashes serve as the resonating bodies and are supported along their length by two poles. The instrument is hung around the neck and beaten with two small sticks . . . Most of the time four marimbas are played . . . When all the instruments play together a really harmonious effect is created from a distance. From close up one can hear the clatter of the sticks, which makes a lot of noise.6
The clatter described was probably not due to the movement of the sticks. It is very likely that small mirlitons were the reason. They were fixed inside the resonating bodies and started vibrating when the instrument was being played – a device still used today for calabash resonaters. The resulting rattling noise is in accordance with the widespread ideal of sound in Africa, where the notes are distorted or alienated. That is why drums are equipped with snares or small rattles, or why no bridge is used when making lyres, so that the strings run directly over the leather top cover of the resonating body which then vibrates, too; lamellophones sometimes have small iron rings attached to the lamellas. Merolla therefore seems to document a musical principle that is important to the current music of many African cultures. Similar significant evidence can be found in the reports by the legendary Dominican father Jean-Baptiste Labat who lived on the French Antilles from 1694 to 1705 and published an extensive travelogue in 1722. This report was translated into German at the end of the 18th century. In 1984 Heinrich Pleticha edited and published a new abridgement.7 Lavat was one of the first to describe the daily life of African slaves, how they dealt with their suffering, and their strategies for coping with their degrading situation. Plenty of space is given to descriptions of the cul5 6 7
Cf. ibid. Girolamo Merolla. Breve, e svccinta Relatione del Viaggio nel regno die Congo nell’ Africa Meridionale. Naples, 1687. 170-171. Jean-Baptiste Labat. Pater Labats Sklavenbericht. Abenteuerliche Jahre in der Karibik 1690-1705. Ed. Heinrich Pleticha. Stuttgart: Edition Erdmann, 1984 (French original: Nouveau voyage aux Isles de l’Amérique, contenant l’histoire naturelle de ces pays, l’Origine, les Mœurs, la Religion & le Gouvernement, des Habitans anciens & modernes. Les Guerres & les Evenemens singuliers qui y sont arrivez pendant le long sejour que l’Auteur y a fait. Paris, 1722).
274
Andreas Meyer
tural life. The description of the music, however, shows some disapproval. Nonetheless, Father Labat compares it with varieties of European folk music. Their music does not sound particularly pleasant, and the notes are not quite in tune. However, there are still some people who find this harmony pleasant, just like that of Spanish or Italian peasants, who all have guitars and play them badly. Whether they are right or wrong I shall leave to your judgement.8
Furthermore the father confirms that many slaves have got a good ear, a step which keeps time, and instrumental skills. In this context he describes a popular string instrument: Nearly all of them play a sort of guitar made from a calabash covered with leather which has been scraped like parchment and which has a rather long neck. They are strung with only four stings which they make from horse hair … or dried intestines and which they grease with Palmchristi oil. The strings are lifted up by a bridge, a broad thumb’s width above the skin of the calabash. They play this instrument by plucking the strings and by beating on it.9
It is obviously one of the earliest descriptions of the banjo. Possible models for this type of instrument can be found in various West African regions, for example in the territory of the Hausa in what in now Nigeria. The spike lutes, as they are called, which are common there (which have a tailpiece spiking the resonating body)10 are also equipped with bodies made of calabashes and are partly beaten and partly plucked. But these instruments in West Africa normally only have two strings.11 The banjo with four strings only developed in the New World as a result of transcultural processes – at the very latest in the 17th century, as is shown by Labat’s report. A further source can be found in the descriptions by the Dutch commercial traveler Willem Bosman who traveled along the Coast of Guinea at the end of the 17th century. In 1705 the English translation of his book titled A New and Accurate Description of the Coast of Guinea was published with historical, geographical, and scientific notes on the Ivory, Gold, and Slave Coasts, that is, the southern region of present day Ivory Coast, Ghana, Togo, Benin, and Nigeria.12 He does not describe his trade partners in very flattering terms: 8 9 10 11 12
Ibid., 198. Ibid. Cf. Ulrich Wegner. Afrikanische Saiteninstrumente. Berlin: Museum für Völkerkunde, 1984. 114. Cf. ibid. 258. Willem Bosman. A New and Accurate Description of the Coast of Guinea. Divided
In Sound Similar to the Harps
275
The Negroes are all without exception, Crafty, Villainous and Fraudulent, and very seldom to be trusted: being sure to flip no opportunity of cheating any European, nor indeed one another. A Man of Integrity is as rare among them as a white Falcon, and their Fidelity seldom extends farther than to their Masters; and it would be very surprising if upon a scrutiny into their Lives we should find any of them whose perverse Nature would not break out sometimes; for they indeed seem to be born and bred Villains.13
This negative attitude is continued when describing some musical instruments from the Gold Coast: Their Musical Instruments are various, and very numerous, but all of them yield a horrid and barbarous Shocking Sound. The chief of them are the mentioned Horns, made, as I have already told you, of small Elephant’s Teeth; . . . Their second sort of Instruments are their Drums; of which there are about ten several sorts, but most of them are excavated Trees covered at one end with a Sheep’sskin, and left open at the other; which they set on the Ground like a KettleDrum, and when they remove it they hang it by a String about their Necks. They beat on these Drums with two long Sticks made Hammer-fashion, and sometimes with a straight Stick or their bare Hands; all which ways they produce a dismal and horrid Noise. The Drums being generally in consort with the blowing of the Horns, which afford the most charming Asses Musick that can be imagined: to help out this they always get a little Boy to strike upon a hollow piece Iron with a piece of Wood; which along makes a Noise more detestable than the Drums and Horns together.14
The hollow piece of iron which he refers to and which he says sounds louder than the drums and horns together is actually an iron bell. These instruments play an important part in the percussion music in southern Ghana (the former Gold Coast and source of the report) and also in many other sub-Saharan regions of Africa, as what are called “timeline patterns” are played on them, rhythmic phrases which are repeated cyclically and which serve the other musicians as an orientation-point.15 Each percussionist plays his rhythm by relating it to the timeline pattern. Bosman’s report is apparently an early testament to this basic principle of ensemble playing in Africa. When talking about drums with resonating bodies made from tree trunks that are played with hammer-like sticks, Bosman is obviously thinking of big, representative instruments. Among the Akan peoples in southern and central Ghana as well as the Ewe in eastern Ghana and
13 14 15
into the Gold, the Slave, and the Ivory Coast. London: Knapton, 1705 (Dutch original: Nauwkeri Beschrving von de Guinese goudtand en slaven-Kust. Utrecht, 1704). Ibid. 117. Ibid. 138f. Cf. Gerhard Kubik. Theory of African Music. 2 vols. Wilhelmshaven: Noetzel, 1994. Vol. 1, 44f.
276
Andreas Meyer
Fig. 4: Ensemble with Atumpan drums from Asante, Ghana. The Atumpan drums lying in a frame are played by the leader of the ensemble with two hook-shaped beaters.
Togo, instruments of considerable size (height 31-55 inches) are known which are still used in rituals today and which are always played with hook-like sticks. In the Asante region of Ghana they are called Atumpan and Bomaa16 (fig. 4). Other instruments that are customary today among the Akan, for example drums in the shape of an hourglass or percussion instruments made from calabashes, originate from the Gur peoples who live in northern Ghana, in Burkina Faso, and in neighbouring regions. Many musicians in Ghana are of the opinion that the big drums also originated there and then reached the southern regions.17 If this type was already at home in the coastal area in the 17th century though, then this assumption seems to be rather speculative if not wrong. The typical instruments from the North are not mentioned by Bosman and may therefore have reached the southern regions of Ghana only at a later stage. The sources have in common the idea that most of the types of musical instruments described can be clearly assigned to definite cultures 16 17
Andreas Meyer. Afrikanische Trommeln. West- und Zentralafrika. Berlin: Museum für Völkerkunde, 1997. 55-68. Personal communication during my field trips to Ghana in 1993, 1997, and 2000.
In Sound Similar to the Harps
277
or cultural regions. Many are still in use today. The same is true of the principles of music making described: the use of timeline patterns as a metrical base and the ways of creating additional vibration when instruments are played. This obvious continuity corresponds with an idea widely expressed in older ethnomusicological texts according to which African musical cultures follow an unchanged tradition “often seemingly without history.”18 In reality though, African music has always been in a process of change due to social, economic, political, and individual artistic factors. Like all music, it is characterized equally by stability and change. In Ghana for example, as was mentioned above, cultural contacts between different ethnic groups were the reason for the dynamic development of the music. Recent research shows that the history of the Mbira music of the Shona in Zimbabwe indicates a high measure of enthusiastic experimentation by the instrumental players. It is apparent that in precolonial times a system developed from a simple sequence of notes played simultaneously that had various means of modulation within modal structures.19 The musical development goes hand in hand with the morphology of the modern Mbira instruments, comparatively more complex than in Santos’s descriptions. The language and tone of the sources are very different from each other. Bosman’s harsh tone is especially striking. Unlike Santos, Merolla, Labat, and Praetorius, Bosman was a profane trader with profane interests who was also involved in the slave trade and therefore could hardly find the time and leisure for cultural values. The alien character of the culture of his trade partners certainly hindered his economic intentions. Morevover, the music with which he was confronted was stylistically different from the lamellophone described by João dos Santos, which has quite a warm and pleasant sound in the European sense, when referring to actual ways of playing. African percussion music must have sounded unfamiliar to Bosman and not pleasant, especially when side blown horns were added. There was no chance that a “harmonious effect” (Morella), such as that a European would have been used to, could be achieved. It can be assumed that Santos and Morella looked more thoroughly into the music. To be successful in their mission the clergymen had to adapt to some extent to village life and therefore absorb the forms of 18 19
Josef Kuckertz. “Einleitung.” Außereuropäische Musik in Einzeldarstellungen. Ed. Edition MGG. Kassel, Basel, London: Bärenreiter Verlag, 1980. 22. Klaus-Peter Brenner. Chipendani und Mbira. Musikinstrumente, nicht-begriffliche Mathematik und die Evolution der harmonischen Progression in der Musik der Shona in Zimbabwe. Göttingen: Vandenhoeck und Ruprecht, 1997. 359-74.
278
Andreas Meyer
cultural expression without prejudice. But Bosman’s disparaging remarks read like the harbingers of the 19th century, the Age of Imperialism, with a language that is an expression of Europeans’ feelings of superiority as justification for the economic dispossession of the natives. “Horrible noise” and “hellish instrumental music” are the attributes with which, for example, the geographer Gerhard Rohlf describes the ritual music of the Kanuri in present-day Nigeria after crossing the Sahara in 1865.20 A similar tone is used by Eduard Hanslick, the famous musical critic and founder of musicology in Vienna, when he characterizes nonEuropean music in his aesthetic text “Vom Musikalisch Schönen” (“On the Musically Beautiful”) which was published for the first time 1854: When the South Sea Islander bangs rhythmically with bits of metal and wooden staves and along with it sets up an unintelligible wailing, this is the natural kind of “music,” yet it just is not music. But what we hear a Tyrolean peasant singing, into which seemingly no trace of art penetrates, is artistic music through and through. Of course, the peasant thinks that he is singing off the top of his head. For that to be possible, however, requires centuries of germination.21
This way of thinking, which maintains that contemporary foreign musical cultures are equivalent to past stages of a development which European music has already gone through and left behind, was taken up at the beginning of the 20th century by the early representatives of comparative musicology. Only the ethnomusicology that developed after Second World War, as a result of a long and laborious process, could finally disprove this idea. It was a real achievement to recognize that the music of each ethnic group in itself represents “a sophisticated and complex organism.”22 In the 17th century one was obviously closer to this knowledge than in later times. A parallel to this changing perception of African cultures can be found in the musical theater and its depiction of African people on the opera stage. In several operas of the late 18th and early 19th century the black person is always shown as the stupid and ugly person and the foreign element which is made a laughing stock of everywhere. Baroque operas give a different impression, like for example the opera Il color fà la regina by the Venetian composer Carlo Francesco Pollarolo, which 20
21
22
Gerhard Rohlfs. Quer durch Afrika. Die Erstdurchquerung der Sahara vom Mittelmeer zum Golf von Guinea 1865-1867. Ed. Herbert Gussenbauer. Stuttgart: Edition Erdmann, 1984. 138, 248. Eduard Hanslick. On the Musically Beautiful: A Contribution towards the Revision of the Aesthetics of Music. Ed. and trans. Geoffrey Payzant. Indianapolis: Hackett, 1986. 69f. Bruno Nettl. ”Ethnomusicology today.” The World of Music 17,4 (1975): 14.
In Sound Similar to the Harps
279
was first staged in 1700. The opera is set in a fantasy country called “Cambodia.” There are Chinese and African people in the play, there is repeatedly talk of the color of their skin but never in a judgmental context. All the main characters are “di bel sembiante,” of beautiful appearance.23 WORKS CITED Bosman, Willem. A New and Accurate Description of the Coast of Guinea. Divided into the Gold, the Slave, and the Ivory Coast. London: Knapton, 1705. Brenner, Klaus-Peter. Chipendani und Mbira. Musikinstrumente, nicht-begriffliche Mathematik und die Evolution der harmonischen Progression in der Musik der Shona in Zimbabwe. Göttingen: Vandenhoeck und Ruprecht, 1997. Hanslick, Eduard. On the Musically Beautiful. A Contribution towards the Revision of the Aesthetics of Music. Ed. and trans. Geoffrey Payzant. Indianapolis: Hackett, 1986. Hirschberg, Walter. “Early Historical Illustrations of West and Central African Music.” African Music 4.3 (1969): 6-18. Kubik, Gerhard. Theory of African Music. 2. vols. Wilhelmshaven: Noetzel, 1994. Kubik, Gerhard. Kalimba, Nsansa, Mbira – Lamellophone in Afrika. Berlin: Museum für Völkerkunde, 1998. Kuckertz, Josef. “Einleitung“. Außereuropäische Musik in Einzeldarstellungen. Ed. Edition MGG. Kassel, Basel, London: Bärenreiter Verlag, 1980. Labat, Jean-Bapatiste. Pater Labats Sklavenbericht. Abenteuerliche Jahre in der Karibik 1690-1705. Ed. Heinrich Pleticha. Stuttgart: Edition Erdmann, 1984. Läpple, Christian. “Ghanaische Musik im Spiegel der Literatur des 19. und 20. Jahrhunderts.” Weltmusik II. Ed. Peter Ausländer and Johannes Fritsch. Cologne: Feedback Studio Verlag, 1983. 3-18. Meyer, Andreas. Afrikanische Trommeln. West- und Zentralafrika. Berlin: Museum für Völkerkunde, 1997. Merolla, Girolamo. Breve, e svccinta Relatione del Viaggio nel regno die Congo nell’ Africa Meridionale. Naples, 1687. Nettl, Bruno. “Ethnomusicology today.” The World of Music 17,4 (1975): 11-15. Pilipczuk, Alexander. Elfenbeinhörner im sakralen Königtum Schwarzafrikas. Bonn: Verlag für Systematische Musikwissenschaft, 1985. Praetorius, Michael. Syntagma musicum. Vol. II. De organographia. Wolfenbüttel, 1620. Rohlfs, Gerhard. Quer durch Afrika. Die Erstdurchquerung der Sahara vom Mittelmeer zum Golf von Guinea 1865-1867. Ed. Herbert Gussenbauer. Stuttgart: Edition Erdmann, 1984. Schulze, Hendrik. “Afrikaner auf der europäischen Opernbühne des 17. und 18. Jahrhunderts.” A Global View of Mozart. Ed. Generalsekretariat “Mozart 2006 Salzburg.” Salzburg, 2002. 8-11. Santos, Frei João dos. Ethiopia Orienta [1609]. Lisbon: 1891. Wegner, Ulrich. Afrikanische Saiteninstrumente. Berlin: Museum für Völkerkunde, 1984. 23
Hendrik Schulze. “Afrikaner auf der europäischen Opernbühne des 17. und 18. Jahrhunderts.” A Global View of Mozart. Ed. Generalsekretariat “Mozart 2006 Salzburg.” Salzburg, 2002. 8-11.
H. OTTO SIBUM
Machines, Bats, and Scholars: Experimental Knowledge in the Late Eighteenth and Nineteenth Centuries Some time in the 1830s, Moritz Hermann Jacobi began the construction of a spectacular machine (fig. 1). In the photograph, eight horseshoes can be made out protruding symmetrically from a wooden frame. Opposite these, on a movable disc, are a further eight that can rotate in front of the fixed ones. Attached to the same rotational axis is a contact, the “Commutator,” which turns by means of magnets and reverses their polarization. When the horseshoes are magnetized by an electric current, the unlike poles are attracted, and set the axis of rotation into motion until these are opposite one another. In this position, the “Commutator” reverses the polarization so that the movable magnets are now given the opposite magnetization and are repelled. In this way, a continual circular motion arises. With this, Jacobi thought that he had achieved his big breakthrough as an experimental physicist. However, his “physical perpetuum mobile” as he called it, long remained the subject of dispute, the history of which I would like to sketch out here.1 I. Models Electro-magnetic phenomena have been a challenge for mechanics since they were first investigated, and it is therefore not surprising that they were considered to be manifestations of a so-called fifth element.2 Of the many attempts at establishing this new force as an ontologically primary force, I will name here only the efforts of those experimental natural philosophers who wanted to trace the movements of the planets 1 2
For a more detailed presentation, see H. Otto Sibum. “Experimentalists in the Republic of Letters.” Science in Context 16 (2003): 89-120. Cf. Jörg Meya and Heinz Otto Sibum. Das fünfte Element. Wirkungen und Deutungen der Elektrizität. Reinbek b. Hamburg: Rowohlt, 1987.
Machines, Bats, and Scholars
281
Fig. 1: Reproduction of the electro-magnetic machine, University of Oldenburg.
back to electrical interactions. In the seventeenth century, dynamic models of the movements of the planets were powered, for the most part, by precise mechanical clock mechanisms, simultaneously confirming the idea of the world as a clock. However, with the electrical research of the eighteenth century we see the first attempts to imitate the movement of the planets using electrical point discharge. These “electrical orreys” were mostly developed in England. In Germany, spectacular celestial phenomena such as the Northern Lights – aurora borealis – were modeled using artificially created electrical discharge phenomena in glass tubes.3 Imitations of nature using electrical experiments not only promised enlightenment about macroscopic processes, but also lead to microscopic discoveries such as electrical current. The invention of the electrical battery goes back, in large part, to Alessandro Volta (fig. 2). Volta called a column composed of pairs of plates of copper and zinc separated by small pieces of damp leather the “first artificial electrical organ.” This was an allusion to numerous attempts by experimental phi3
On the electrical planetarium (“electrical orrey”) by George Adams, cf. Alan Morton and Jane Wess. Public and Private Science. Oxford: Oxford University Press, 1993.
H. Otto Sibum Voltasäule: © Musée des arts et métiers – CNAM, Paris
282
Fig. 2: Anatomical cross-section of a torpedo fish and electrical column by Volta.
losophers to create a model of the electrical organ of the torpedo fish. In the eighteenth century, it was widely known that a few exotic fish such as the torpedo fish or the electric eel (gymnotus electricus) produced, on contact, painful effects similar to an electric shock. Henry Cavendish built a model of the torpedo fish in order to study these electrical effects in detail. Anatomical drawings of the fish and wax models of the electrical organ also served the Italian physicist Volta as a model for the construction of his electrical column. The part of the electrical organ designated H consists of a column of parallel muscle fibers; the gaps between contain fluid. One easily recognizes that the Volta column on the right is assembled from parallel copper and zinc plates, each separated by pieces of damp leather. Importantly, however, Volta, after long experimentation, eventually insisted that this electricity had nothing to do with the “animal electricity” identified by Luigi Galvani in dissected frogs, but that only the contact between metals such as copper and zinc was essential. Without being able to go into these inquiries in detail here, we can see that these attempts at modeling led to important insights concerning electricity and magnetism, which in the eighteenth century, however, could not initially be integrated into the canon of physical science. Only towards the end of the century do we notice several important points of
Machines, Bats, and Scholars
283
interaction between traditional, scholarly natural science and the new experimental philosophy.4 In the eighteenth century it was the adherents of Newtonian mechanics who set the tone for research into electricity and magnetism, by attempting to develop this as the universal frame of explanation for electrical and magnetic phenomena. One of the best known figures in this connection was the French engineer Charles-Augustin de Coulomb. With a torsion balance especially constructed for the purpose, he demonstrated to the French public the validity of the Newtonian law of force for electrical and magnetic phenomena. The repetition of this experiment showed, however, that the Newtonian law could not be clearly demonstrated with this experiment. For this reason, until 1800, outside France, no appreciable reception of the experiment could be observed.5 It is not surprising, therefore, that Jacobi did not follow the French doctrine without criticism, instead giving priority to theoretical insights achieved using the electro-magnetic model. In 1835, he wrote: After my many experiments, I can now announce to the public that magnetism as a force like gravity acts as a pure function of space . . . The instantaneous reversal of the poles leads [moreover] to an infinitely accelerated movement.
With this, the construction of a physical perpetuum mobile was achieved. A mechanical perpetuum mobile is not possible because a moving force can only produce an effect equal to itself; that there might be a physical one is certainly possible since it would require a force [Triebkraft] that could be randomly produced by mere movement.6
On the one hand, this fine distinction takes into account the scientific notion that there could be no machines whose mechanical effect could be greater than what is necessary for powering them. On the other hand it enabled researchers to examine new types of natural force such as 4
5
6
Cf. on Henry Cavendish, “An Account of Some Attempts to Imitate the Effects of the Torpedo by Electricity.” Philosophical Transactions of the Royal Society 66 (1776): 196-225; on Alessandro Volta and his “first artificial electrical organ,” cf. Meya and Sibum. Fünfte Element. 132-41 as well as Giuliano Pancaldi. Volta. Science and Culture in the Age of Enlightenment. Princeton: Princeton University Press, 2003. Christine Blondel and Mathias Dörries, eds. Restaging Coulomb. Usages, controverses et réplications autour de la balance de torsion. Florence: L.S. Olschki, 1994; H. Otto Sibum. “Charles-Augustin Coulomb (1736-1806).” Die Großen Physiker. Von Aristoteles bis Kelvin. Ed. Karl von Meyenn. Munich: Beck, 1997. 243-62. Moritz Hermann Jacobi. “Über die Benutzung der Naturkräfte zu menschlichen Arbeiten.” Vorträge aus dem Gebiete der Naturwissenschaften und der Oekonomie gehalten vor einem Kreise gebildeter Zuhörer in der physikalisch-ökonomischen Gesellschaft zu Königsberg. Ed. Karl Ernst von Baer. Königsberg, 1834. 105.
284
H. Otto Sibum
electricity and magnetism in order to verify how far these were in a position to produce sustained movement. From this scientific perspective, even a normal barometer could be treated as a physical perpetuum mobile, since the natural fluctuations in air pressure result in a constant movement in the display device.7 Jacobi’s electro-magnetic machine model now showed that the “constant circulation of things,” seen for example in the orbit of the planets, was not necessarily of a mechanical, but rather a magnetic nature. Jacobi’s inventiveness had conceived a machine, which was in a position to wrest from nature an automatic continual movement by cunning. For him, his discovery meant a “worldhistorical event,” as he believed he had shown that for a deeper understanding of nature, it was no longer clockmakers, but now those investigating electricity who would be called upon to investigate its potential. II. The Motor of Civilization For Jacobi, however, the electro-magnetic machine was also the motor of civilization. The term motor is a scholarly neologism of the early nineteenth century. Borrowing from the Latin verb ‘movere’ = to move, the word “motor” was used in the same way as “engine” or also “force.” The electromotor is thereby of great significance for the understanding of the historical development of the man-machine relationship. Jacobi’s model represents a crucial example in the historical transformation of the conceptions of work and the role of the human body. In 1834, Jacobi commented on this transformation as follows: The physical force of the human race – and this is a striking peculiarity of our time – we see made use of increasingly rarely; a higher intelligence announces its intention to grasp the raw forces of nature in more determined way, and to constrain them to a certain purpose.8
As a civil engineer, Jacobi was naturally well acquainted with British steam technology and its implications for the transformation of work practices. But only after examining general questions of civil engineering, philosophy, and the emerging science of anthropology did he formulate a model of civilization that considered the notion of work and the machine as central quantities. In an unpublished lecture manuscript 7
8
The difference between “mechanical perpetuum mobile” and “physical perpetuum mobile” is presented at length in Gehler’s Physikalischem Lexikon. Cf. the entry “Perpetuum Mobile” in D. Johann Samuel Traugott Gehler. Physikalisches Wörterbuch. 11 vols. Leipzig, 1833, vol. 7, 408-23. Jacobi. “Über die Benutzung der Naturkräfte.” 101.
Machines, Bats, and Scholars
285
he emphasizes how privileged he has been to be able to experience at first hand how each art creates a “second nature, a nature shaped according to human motives and for human purposes.” Therefore, for an understanding of the development of our civilization, it was not the pairing of artificial and natural that was crucial, but the confrontation of spirit and nature: The human spirit, whose necessary condition is moral development, does not find, nor should it, its complete gratification in nature; the divine spirit living within it demands that it freely and independently build its own world, for whose creation it uses nature. But since it is already beyond nature, it achieves this in reality only by conquering it. This, however, is only possible by means of work. The quantity of industriousness, intellectual or biblical, thus also provides a measure of moral development.9
Here Jacobi already presents himself – one could say – as the bourgeois philosopher of work, as someone who has recognized industry – industria – as the motor and moral economy of civilization. But as the result of machines, precisely this industry experienced a fundamental transformation, presenting a great challenge to science. And so he continued: When liberation from nature by work is completed, then the latter in its universal meaning initially attempts if not to do without strict categories, space, and time, then to randomly, i.e. freely, shape their mutual designation . . . Furthermore, man’s desire to make nature into an organ simultaneously contains the principle of the machines, which man places between himself and nature and which one should [firstly] consider as an emancipation from material work, which gives him the ability and the leisure to busy himself with higher things, and finally to reach the realm of the spirit.10
His conception of emancipation from material work by machines corresponds, in important points, to the classical ideas of the British advocates of industrialization. However, Jacobi also criticized the British practice of industrialization based on steam technology, which seemed to subject the transformation of work to an all-controlling spirit of economics, and therefore could be considered responsible for the undermining of craft and related forms of knowledge. Jacobi’s criticism can only be understood in terms of the conception of work that was predominant in Germany in his time. It was precisely this idea that had still not distinguished between work and labor as quantitative measures. Besides, artisanal work was still 9
10
Moritz Hermann Jacobi. “Von der eigentlichen Bedeutung des Luxus und der Mode” (unpublished lecture manuscript). St. Petersburg Academy of Sciences Archive, fond 187/op.1/dele 330. St. Petersburg, undated [ca. 1835]. 2f. Ibid. 3f.
286
H. Otto Sibum
seen as a process of integrating intellectual and manual labor. If industry should be treated as a motor of civilization, according to Jacobi, it was the responsibility of science and education to manage the shaping of this second nature without destroying the organic link between man and nature. III. Handwerksgelehrte The machine, considered in the field of tension between the mechanical arts and the sciences, leads us to the last and most significant part of this history of knowledge. Since the early modern period, the scholarly opinion of the “art of experimenting” has varied from the total negation of its epistemological value, to the conviction dominating the nineteenth century that this form of research is the only way to properly grasp the origins of natural phenomena. One of the disputes underlying this controversy was that the physical manipulation of objects characterizing the experiment was not derived from the proven practice of the traditional scholars [Schriftgelehrten]. Quite the contrary – according to their self-understanding, knowledge and praxis were clearly distinct. Even enlightened philosophers such as Denis Diderot, who in fact described the mechanical arts as a form of knowledge, conceded that this knowledge lay outside enlightened discourse. Diderot’s encyclopaedia project represented an attempt to provide a language for the knowledge of the practitioner that could be understood by anyone. With all due respect to this extraordinary achievement, we must nevertheless concede that this literary rapprochement also reduced complex forms of the knowledge implicit in praxis to visual representations or descriptions of manual techniques, and thereby perpetuated the traditional division between epistemology and praxis. In order to bridge the divide between traditional science and the mechanical arts, between theory and praxis, the engineer, since the middle of the eighteenth century, has been seen as the ideal mediator – as the third man. In this context, the philosopher Christian Wolff wrote: In such circumstances, a third man would be needed, who could in himself unite science and art, in order to correct the theorists’ infirmities and to combat the prejudice of the lovers of the arts, as if they could be therein complete without the theory, and leave it [theory] to the idle heads good-for-nothing in the world. . . . Hence . . . he [Leupold, author’s note] compared himself to a bat, tolerated among neither birds nor quadrupeds, and he complained that he was hated by the practitioners of art as well as despised by the theorists, for he wanted by his nature to be celebrated as a remarkable man by both, and to share fame in the learned world with the latter and happiness at court with the former.11 11
Christian Wolff. “Preface.” Bernard Forest de Belidor. Architectura Hydraulica.
Machines, Bats, and Scholars
287
In the historical process of the integration of the physica experimentalis within the ‘Republic of Letters’ [Gelehrtenrepublik] emerging here, the experimentalists experienced at first hand the advantages and disadvantages of the hopeful new position of the “Third Man.” As it seems, this was as difficult to categorize as bats. In addition, it took traditional science to the limits of its self-understanding. It might therefore be asked whether these studies of the experimentalists, uniting intellectual and manual labor, are a matter of a specific form of knowledge [Wissen]. Can this be described as science [Wissenschaft]? The answer would be very different depending on which understanding of knowledge in general, and scientific knowledge in particular was being favored. The process of differentiation between experimental knowledge and science seen here is therefore the expression and integral component of a largely still unwritten historical process that is commonly described as the second scientific revolution. Until now, this very heterogeneous field of practices for the production of knowledge has not been completely exploited, let alone worked through in all its individual facets. Jacobi’s electro-magnetic machine is therefore only one piece of the mosaic in the process of uncovering the texture of this complex scientific transformation around 1800 – less a scientification of applied knowledge, than a differentiation between knowledge and science. Therefore, if one wants to measure the changing status of experimental science in this historical period, it is helpful to call on the prevalent scholarly position. The understanding, dominating the republic of letters, of scientific knowledge as universal, autonomous, and permanent, was closely bound up with the hegemony of the written text. As early as in the middle of the eighteenth century, many generations of natural philosophers were required to remove the eipstemological stigma from the art of the experiment in order to be able to position the knowledge implicit in the praxis of the experimenters within the republic of letters. The quickly expanding experimental investigation of electricity and magnetism played a key role in changing the scholarly opinion of the epistemological status of the experiment. The unknown effects produced daily in these experiments demonstrated the untenable nature of the traditional and, in the first half of the eighteenth century, generally accepted “scholarly” position that experience produced by the human hand did not deliver new physical truths: Oder: Die Kunst, das Gewässer zu denen verschiedentlichen Nothwendigkeiten des menschlichen Lebens zu leiten, in die Höhe zu bringen, und vortheilhaftig anzuwenden. Augsburg, 1764. 2.
288
H. Otto Sibum
Experience gained in physics through the senses is of a twofold kind: one sort we take from God’s creatures, from fire, air, water, earth, from the stars, flowers etc. the other we gain from artificial things, which are made by human hands . . . But we have no cause to make a great show of it, as if one could discover new and hitherto unknown physical truths through them [artificial things].12
In the course of these researches the latter position became obsolete, since in fact “new and previously unknown physical truths” were established by experiment, and precisely these supported the young experimental physics in its emancipation from natural history. As early as in 1755, the translation of the inaugural lecture of the French experimenter Jean-Antoine Nollet was commissioned to propagate the “physique expérimental” within the republic of letters.13 Nollet insisted that the praxis-oriented method of research should be clearly distinguished from the practices of natural history. The latter is a form of investigation that “manufactures the catalogue of our riches,” but could not investigate the “causes of what happens in the natural world.”14 For indeed, he who endeavours to investigate nature without understanding its history speaks at random and about things that he does not know in the least; but he who knows nothing else of nature than its history justly deserves a place among those natural philosophers who exercise their memory only [Gedächtnisgelehrte]. Accordingly, to practice experimental physics is nothing else than to investigate nature, not only with regards to its effects, but equally with the intention [of studying] the tools by which [nature’s] effects are produced; in short it means to study what [nature] does, in order to be in a position to say how she does it.15
Despite this manifesto, it cannot be denied that until the early nineteenth century, the scholar was above all a writer, who largely produced translations as well as publishing textbooks and compendiums. It is not by chance that the man who most successfully ushered in experimental physics at Göttingen University, Georg Christoph Lichtenberg, was both a writer and an experimenter. The edition of Erxleben’s “Anfangsgründe der Naturlehre,” produced by Lichtenberg, shows very clearly how the “new and previously unknown physical truths” based on the 12
13
14 15
Johann Georg Walch. “Experimental-Physic.” Philosophisches Lexicon . . . Ed. idem. Leipzig, 1733. Quoted in Hans Schimank. “Zur Geschichte der Physik an der Universität Göttingen vor Wilhelm Weber (1734-1830).” Rete. Strukturgeschichte der Naturwissenschaften 2 (1974): 213. Jean-Antoine Nollet. Rede von der nötigen Geschicklichkeit zur Erforschung der Natur, welche er den 15. Mai 1753 bei dem Antritte seines öffentlichen Lehramtes in dem Navarrischen Collegio gehalten. Erfurt, 1755. Ibid. 11f. Ibid. 12.
Machines, Bats, and Scholars
289
experiment, constructively changed the traditional scholarly textual exegesis. The massive invasion of the Lichtenbergian footnotes, to be seen clearly in the sixth edition, shows how quickly the natural knowledge acquired from the physica experimentalis needed to be changed, and the universal truths in the scholarly book of nature needed to be modified.16 Experimental knowledge changed the publication and research practices of the university scholars. Traditional encyclopedic mentalities yielded to an attitude that at the time was aptly described as “scientific invention.” The increasing investigation of nature’s operations, using instruments especially developed for this purpose, led to violent controversies about whether the experimental findings were an invention of a virtuoso instrument maker, or whether it was in fact a question of a discovery.17 The above-mentioned controversy around the Coulombian experiments to determine the quadratic force law of electrical charges shows the problem very clearly. Do the measurement results really prove the validity of the Newtonian action-at-the-distance law for electrical and magnetic forces? Or – as other scholars argued – did just the geometry of the apparatus determine these measurements and, therefore, make of the law a mere artifact? While it would still take decades for this law to achieve universal validity, for the engineer, the instrumental procedure was epistemologically unproblematic. Since the engineer was “natural, as was the countryside which he confronted. His ‘genius’ wrote under the dictation of nature, and it was this same nature that he had to try at all times to transform.”18 Many protagonists of the physica experimentalis shared this view of their research method, and the Prussian reformers of the early nineteenth century specially coined the phrase “scientific invention”, so that this engineer-like invention would have academic credibility within the community of scholars. A further key problem for the integration of experimental knowledge within the traditional scholarly world was the practice of building models. As we have seen, central discoveries in the investigation of electricity and magnetism rested on attempts at modeling, which however 16 17
18
Johann Christian Polykarp Erxleben. Anfangsgründe der Naturlehre. 6th ed. Göttingen, 1794. In the eighteenth century, invention and discovery were frequently still used synonymously. Only around the turn of the century is it possible to observe a polarizing differentiation of these terms. In relation to this, cf. Lorraine Daston. “Introduction.” The Biography of Scientific Objects. Chicago: University of Chicago Press, 2000. Antoine Picon. French Architects and Engineers in the Age of Enlightenment. Trans. Martin Thom. Cambridge: Cambridge University Press, 1992. 229.
290
H. Otto Sibum
have general epistemological meaning, since ultimately every experimental system represents a model. As studies on the practices of modeling in England during this period confirm, it was precisely the different possible solutions in the process of generalization, from the individual manipulable model up to the real object, that determined the legitimacy of the model maker’s knowledge.19 In the course of integrating traditions of experimental knowledge into academic culture, models acted as mediators between the complex forms of knowledge of the practitioners, though they were also denigrated as scientifically meaningless if the problem of true to scale rendering was not convincingly resolved. At the end of the eighteenth century a series of knowledge traditions, those belonging to artisans, instrument makers, and engineers, found themselves in conflict concerning the true meaning and scope. Jacobi’s electro-magnetic machine, but in general all models, were the immediate expression of a new experimental culture, which distanced itself from the pure mathematical tradition, and emancipated itself from craft and the mechanical arts. The public, however, still needed to be convinced that the experimental knowledge thus acquired would have significant consequences for science and society. But what public? Scholars of the caliber of his brother, the mathematician Carl Gustav Jacob Jacobi, considered M.H. Jacobi’s work, with all due sympathy for the art of experimenting, as applied mathematics. Prussian reformers with their gaze fixed firmly on British industrialization seemed to judge his model exclusively in relation to technical utility, and viewed it as a utopia. And it was his brother again who named the core problem: Just recently I read in relation to the E.M. machines in a philosophical Magazine the warning to be on one’s guard against any conclusion about real machines from models, and therefore I will have no peace, and cannot recognize their existence until they really exist, and are no longer handed around with the tea.20
The suggestion that his machine was nothing more than one of the many philosophical toys presented at court or the academies for educated entertainment revealed a widely held scholarly view that in many cases the universal validity of the functional connections of nature demonstrated in a model construction or an experiment had not been prov19
20
Simon Schaffer. “Fish and Ships. Models in the Age of Reason.” Models. The Third Dimension of Science. Ed. Soraya de Chadarevian and Nick Hopwood. Stanford: Stanford University Press, 2004. 71-105. “Carl Gustav Jacob Jacobi an Moritz Hermann Jacobi, 9. Juni 1838”. Briefwechsel zwischen Carl Gustav Jacob Jacobi und Moritz Hermann Jacobi. Ed. Wilhelm Ahrens. Leipzig: Teubner, 1907. 54f.
Machines, Bats, and Scholars
291
en. To a certain degree, Moritz shared this opinion. However, as an experienced experimenter, he held the position that, according to the nature of the thing, only previous experience could serve as a guideline for the construction of a machine. He welcomed the fact that “these principles to be investigated did not merely have a technical side, but deliver meaningful scientific results and throw light on the nature of these enigmatic agents.” Research into the machine would therefore serve in large part to determine of the laws at the basis of electro-magnetic phenomena. His insistence on using experience acquired from model construction as a guideline for further research, instead of an over-hasty reversion to established theories, points to a specific practice of the construction of theory that found further extension among experimenters. Thus the British mathematician and civil servant, George Atwood, described this practice of theorizing as one that develops systematic rules from experience and observation alone. It clearly distinguished itself from the academic practice of theorizing that followed the application of the pure rules of mathematics. According to the chemist Justus von Liebig, this latter academic practice seemed to predominate particularly in Germany. Jacobi followed Atwood’s technique. Therefore he did not recognize the behavior of his machine as being describable or determined by the established theory of electrostatics. With his research program he hoped instead to open up a new field of research, and with his research method he indeed managed to identify such central physical phenomena as mutual induction. But finally he also bowed to the pressure of his patron, the Tsar of Russia, and determined the performance of the motor in comparison to steam technology (fig. 3). If this public demonstration of his experimental art was very beneficial for his career in Russia, his main concern was something different: namely, the establishment of experimental knowledge as science. It required massive efforts on both sides to bridge the divide between the scholarly world and the practitioners. He thus wrote to his brother: even today one finds enough influential scholars strongly convinced “that practical work is contrary to the spirit of science and precision.”21 The awarding of an honorary doctorate to Moritz Jacobi in 1835 by the University of Königsberg is therefore undoubtedly a historical turning point. Unquestionably, his brother as well as the physicist Franz Neumann emphatically stood up for this promotion of the experimental sciences. This at first met with considerable resistance, since awarding a doctorate to a virtuoso was not at all self-evident, whether an artist or an experimental scientist. 21
“Moritz Hermann Jacobi an Carl Gustav Jacob Jacobi, 1/13. Februar 1848”. Ibid. 165.
292
H. Otto Sibum
Fig. 3: Contemporary representation of the ‘electro-magnetic boat ride’ by M.H. Jacobi.
The circumstances known to us of the honorary doctorate awarded by the University of Königsberg to the pianist Franz Liszt in 1842 show this clearly. At first, the dean of the philosophy faculty could not reconcile himself to the suggestion, and it took considerable effort to move him to be part of a unanimous resolution. Finally he agreed with the words: “All right, why shouldn’t we award a musician an honorary doctorate. At a time when it is even customary to award a doctorate to chemists.” On the other hand, it was manual work and technical skill that was initially judged detrimental to the spirit of science. And again it was the mathematician Jacobi who contradicted this view. In his eulogy for Liszt, he clearly expressed under what conditions manual skill in the use of instruments, i.e. virtuosity, could be considered a mark of genius: The wonders of technology are only a moment, a means and organ for you to express your higher states of soul. The true master gives us a new revelation of art; he thereby joins the community and the circle of free spirits, who are called upon to represent their time.22
The scholars of the nineteenth century only gradually accepted that this type of virtuosity amongst experimenters could lead to important natural discoveries. Jacobi attempted to give this new thought the most comprehensive expression possible. Therefore he also described his workshop as a mechanical studio, a hybrid of mechanical and liberal arts, to 22
Carl Gustav Jacob Jacobi. “Laudatio für Franz Liszt”. Wissenschaftliche Monatsblätter IV (1842): 176.
Machines, Bats, and Scholars
293
distinguish it from a craftman’s workshop. However, his life also shows very clearly what difficulties the third man faced in overcoming the gulf between theory and praxis. He personally followed a double strategy. On the one hand he continuously attempted to liberate manual activities from their poor reputation by presenting the significance of sensory experience for the development of new theoretical knowledge. On the other hand, mindful of the epistemological and moral standards of the traditional republic of letters, he attempted a permanent demarcation of his experimental knowledge from purely artisanal or even engineering knowledge. In this way the civil engineer was promoted to an artisanscholar. The German term ‘artisan-scholar’ [Handwerksgelehrter] was coined in the second half of the nineteenth century and well reflects the fusing of the experimentalists with the traditional academic elite. What in the eighteenth century were seen as distinct traditions of knowledge, such as for example the mechanical arts and erudite culture [Schriftgelehrtentum], were amalgamated in the nineteenth century into a firmly established group of experimental scientists.23 IV. Conclusion To sum up, it can be remarked that Jacobi’s electro-magnetic machine could only be constituted as a result of the interaction of very different traditions of knowledge which came together in the early nineteenth century. The various modes of looking at this scientific object give expression to a transformation of scientific culture, where the art of experimention still needed to be established as a scientific discipline. Until the nineteenth century, the knowledge of natural phenomena acquired by the construction of machines remained a controversial form of knowledge that still had to be established in academic science, just as during this period the new scientific persona, the experimental scientist, achieved his academic and cultural recognition. Jacobi’s work on his machine, however, also marked a turning point in the instrumental relationship of the scientist to nature determined by experimenting since the early modern period. The invention and application of self-moving machines, and self-registering instruments thus lead to various considerations about the meaning this type of instrument would have for the development of the scientific cognitive ideal of the nineteenth century. This essay marks 23
A detailed presentation of this process of the integration of the experimental sciences within the European and Northern American universities cannot be dealt with here.
294
H. Otto Sibum
the beginning of the investigation of the meaning of the electro-magnetic machine in this process of development.24
Translation: Benjamin Carter WORKS CITED Ahrens, Wilhelm, ed. Briefwechsel zwischen Carl Gustav Jacob Jacobi und Moritz Hermann Jacobi. Leipzig: Teubner, 1907. Blondel, Christine and Mathias Dörries, eds. Restaging Coulomb. Usages, controverses et réplications autour de la balance de torsion. Florence: L.S. Olschki, 1994. Brain, Robert Michael. The Graphic Method. Inscription, Visualization, and Measurement in Nineteenth-Century Science and Culture. Ph.D. University of California, 1996. Cavendish, Henry. “An Account of Some Attempts to Imitate the Effects of the Torpedo by Electricity.” Philosophical Transactions of the Royal Society 66 (1776): 196-225. Daston, Lorraine and Peter Galison. “The Image of Objectivity.” Representations 40 (1992): 81-128. Daston, Lorraine, ed. The Biography of Scientific Objects. Chicago: University of Chicago Press, 2000. Erxleben, Johann Christian Polykarp. Anfangsgründe der Naturlehre. 6th ed. Göttingen, 1794. Gehler, D. Johann Samuel Traugott. Physikalisches Wörterbuch. 11 vols. Leipzig, 1825-1845. Jacobi, Carl Gustav Jacob. “Laudatio für Franz Liszt.” Wissenschaftliche Monatsblätter IV (1842): 175-76. Jacobi, Moritz Hermann. “Über die Benutzung der Naturkräfte zu menschlichen Arbeiten.” Vorträge aus dem Gebiete der Naturwissenschaften und der Oekonomie gehalten vor einem Kreise gebildeter Zuhörer in der physikalisch-ökonomischen Gesellschaft zu Königsberg. Ed. Karl Ernst von Baer. Königsberg, 1834. 99-123. Jacobi, Moritz Hermann. “Von der eigentlichen Bedeutung des Luxus und der Mode” (unpublished lecture manuscript). St. Petersburg Academy of Sciences Archive, fond 187/op.1/dele 330. St. Petersburg, undated [ca. 1835]. 24
Lorraine Daston and Peter Galison argue that “mechanical objectivity” in the nineteenth century must be seen in close connection with the development of photography and self-registering instruments. Lorraine Daston and Peter Galison. “The Image of Objectivity.” Representations 40 (1992): 81-128. Simon Schaffer has pointed out that the notion of genius prevalent during this period can only be understood in relation to the use of self-registering instruments. Simon Schaffer. “Self Evidence.” Critical Inquiry 18 (1992): 327-62. Robert Brain emphasizes that the self-registering instrument invented by James Watt for the automatic recording of the performance of a steam engine marks the beginning of the graphic method in science and culture. Robert Michael Brain. The Graphic Method. Inscription, Visualization, and Measurement in Nineteenth-Century Science and Culture. Ph.D. University of California, 1996.
Machines, Bats, and Scholars
295
Meya, Jörg and Heinz Otto Sibum. Das fünfte Element. Wirkungen und Deutungen der Elektrizität. Reinbek b. Hamburg: Rowohlt, 1987. Morton, Alan and Jane Wess. Public and Private Science. Oxford: Oxford University Press, 1993. Nollet, Jean Antoine. Rede von der nötigen Geschicklichkeit zur Erforschung der Natur, welche er den 15. Mai 1753 bei dem Antritte seines öffentlichen Lehramtes in dem Navarrischen Collegio gehalten. Erfurt, 1755. Pancaldi, Giuliano. Volta. Science and Culture in the Age of Enlightenment. Princeton: Princeton University Press, 2003. Picon, Antoine. French Architects and Engineers in the Age of Enlightenment. Trans. Martin Thom. Cambridge: Cambridge University Press, 1992. Schimank, Hans. “Zur Geschichte der Physik an der Universität Göttingen vor Wilhelm Weber (1734-1830).” Rete. Strukturgeschichte der Naturwissenschaften 2 (1974): 207-52. Schaffer, Simon. “Self Evidence.” Critical Inquiry 18 (1992): 327-62. Schaffer, Simon. “Fish and Ships. Models in the Age of Reason.” Models. The Third Dimension of Science. Ed. Soraya de Chadarevian and Nick Hopwood. Stanford: Stanford University Press, 2004. 71-105. Sibum, H. Otto. “Charles-Augustin Coulomb (1736-1806).” Die Großen Physiker. Von Aristoteles bis Kelvin. Ed. Karl von Meyenn. Munich: Beck, 1997. 243-62. Sibum, H. Otto. “Experimentalists in the Republic of Letters.” Science in Context 16 (2003): 89-120. Wolff, Christian. “Vorrede.” Bernard Forest de Belidor. Architectura Hydraulica. Oder: Die Kunst, das Gewässer zu denen verschiedentlichen Nothwendigkeiten des menschlichen Lebens zu leiten, in die Höhe zu bringen, und vortheilhaftig anzuwenden. Augsburg, 1764.
PETER GALISON/LORRAINE DASTON
Scientific Coordination as Ethos and Epistemology Introduction: The Many-Headed Knower It is just possible that the young René Descartes once dreamed of a natural philosophy deduced entirely from clear and distinct ideas by a solitary thinker, as Euclid’s Elements were deduced entirely from axioms, definitions, and postulates. But even Descartes had abandoned the hope of a one-man science by the time his Discours de la méthode appeared in 1637: from first principles “I discovered skies, stars, an earth, and even, on the earth, water, air, fire, minerals, and several other things which are the commonest of all,” but once he descended to further particulars, he realized that it would be necessary “to make use of many experiments” – experiments that he realized far exceeded his own time and resources, “were they a thousand times greater than they are.”1 He hoped that wealthy patrons might finance a legion of paid experimenters under his direction. Thereafter, no one, even the most resolute rationalist, ever imagined the science of nature as a single-handed operation. Opinions in the seventeenth century differed as to how much manpower would be needed for how long, and how it should be best organized, from Leibniz’s wildly optimistic view that a team of scholars would need a mere five years in order to translate all human concepts into an arithmetic “characteristica universalis”2 to Francis Bacon’s vision in the New Atlantis of a whole society permanently organized 1
2
“. . . et il me semble que, par la, i’ay trouué des Cieux, des Astres, vne Terre, & mesme, sur la terre, de l’Eau, de l’Air, du Feu, des Minéraux, & quelques autres telles choses, qui sont les plus communes de toutes & les plus simples . . .;” / “. . . en si grand nombre, que ny mes mains, ny mon reuenu, bien que i’en ieusse mille fois plus que ie n’en ay, ne sçauroient suffire pour toutes . . .” René Descartes. “Discours de la méthode” [1637]. Œuvres de Descartes. Ed. Charles Adam and Paul Tannery. Paris: Léopold Cerf, 1902. Vol. 6, 64-65 and 72-73. Gottfried Wilhelm Leibniz. Untitled fragment [1677]. Die philosophischen Schriften von Gottfried Wilhelm Leibniz. Ed. Carl Immanuel Gerhardt. Berlin: Weidmann, 1875-90. Vol. 7, 184-89.
Scientific Coordination as Ethos and Epistemology
297
around a well-staffed and hierarchically organized research institute known as the House of Solomon.3 And since the seventeenth century, the most diverse models of scientific collectivities have been invented and institutionalized, from the Enlightenment Republic of Letters to the late twentieth-century research laboratory in high energy physics or genetics. Yet however various these collectivities, real and imagined, may be, they all address the problem of how knowledge about nature can be, indeed must be, the product of many hands and heads. This is a problem of the division of labor and the multiplication of laborers, akin to any other such problem in the organization of work: how to analyze a complex inquiry into modular parts, find willing and able hands to undertake each part, and efficiently integrate the results. But it is more than a problem in political economy transferred to science. Even if legions of well-trained, specialized scientific workers were to fan out over the length and breadth of nature and send their data to some updated successors of Bacon’s “Interpreters of Nature” to tabulate and synthesize, the challenges facing every scientific collectivity would not be solved. Gathering and sifting data is not enough, even if armies of researchers are enlisted. Some means must be found to identify fields worthy of inquiry, to standardize and coordinate tools of inquiry (be these the eye of the botanist, the thermometers and barometers of the meteorologist, or the tables of the astronomer), and, above all, to plot a further course of investigation: what next? The prize questions of the eighteenth-century academies, the magisterial review articles of the late nineteenth century, the gigantic grant proposals of the early twenty-first are all means to these ends, albeit within quite different scientific collectivities: the Enlightenment Republic of Letters, Victorian Men of Science, Wartime Collaborations, Post-World War II Scientific Communities. Scientific collectivities are historically situated. Unsurprisingly, they borrow and bricolage forms of labor in their ambient societies: early modern scientific endeavors drew heavily on the organization of work in the household, the workshop, and the court;4 mid-nineteenth-century 3 4
Francis Bacon. “New Atlantis” [1627]. Lord Bacon’s Works. Ed. Basil Montagu. London: William Pickering, 1825-34. Vol. 2, 348-50 and 361-79. Owen Hannaway. “Laboratory Design and the Aim of Science. Andreas Libavius versus Tycho Brahe.” Isis 77 (1986): 585-610; Steven Shapin. “The House of Experiment in Seventeenth-Century England.” Isis 79 (1988): 373-404; Londa Schiebinger. The Mind Has No Sex? Women in the Origins of Modern Science. Cambridge and London: Harvard University Press, 1989. 66-96; Bruce T. Moran. The Alchemical World of the German Court. Occult Philosophy and Chemical Medi-
298
Peter Galison/Lorraine Daston
science self-consciously likened itself to the factory;5 mechanization and Taylorism featured prominently in mid-twentieth-century science as well as in coeval industry.6 Somewhat less obviously, they channel forms of cultural authority as well, as in the case of the earliest scientific academies, which sought royal patronage and actively recruited aristocratic members.7 Conversely, scientific collectivities are occasionally held up as models for society at large, as when John Stuart Mill looked to the physical sciences for a peaceable solution to modern conflicts over legitimate authority or when Jacob Bronowski praised the scientific community as an ideal democracy.8 Yet scientific collectivities are seldom simply microcosms of their economic and social macrocosms; they invent new ways of working, communicating, adjudicating, and authorizing. They are obliged to innovate because their aims are, in contrast to ordinary polities, epistemological, as well as economic, social, and political. Hence every scientific collectivity turns the tools of its age to ends other than those for which they were originally forged: the Renaissance court was not primarily designed to probe the secrets of nature, and yet it was ingeniously adapted to those ends;9 the
5
6 7
8
9
cine in the Circle of Moritz of Hessen, 1572-1632. Stuttgart: F. Steiner Verlag, 1991; Mario Biagioli. Galileo, Courtier. The Practice of Science in the Culture of Absolutism. Chicago and London: University of Chicago Press, 1995; Paula Findlen. “Masculine Prerogatives. Gender, Space, and Knowledge in the Early Modern Museum.” The Architecture of Science. Ed. Peter Galison and Emily Thompson. Cambridge and London: MIT Press, 1999. 29-58; Alix Cooper. “The Household.” The Cambridge History of Early Modern Science. Ed. Katharine Park and Lorraine Daston. Cambridge: Cambridge University Press, 2006. 224-37. David Cahan. An Institute for Empire. The Physikalisch-Technische Reichsanstalt, 1871-1918. Cambridge: Cambridge University Press, 1989; Norton Wise and Crosbie Smith. Energy for Empire. A Biographical Study of Lord Kelvin. Cambridge: Cambridge University Press, 1989; Timothy Lenoir. Instituting Science. The Cultural Production of Scientific Disciplines. Stanford: Stanford University Press, 1997. Peter Galison. Image and Logic. A Material Culture of Microphysics. Chicago and London: University of Chicago Press, 1997. Roger Hahn. The Anatomy of a Scientific Institution. The Paris Academy of Sciences, 1666-1803. Berkeley and Los Angeles: University of California Press, 1971; Michael Hunter. The Royal Society and Its Fellows 1660-1700. Morphology of an Early Scientific Institution. Chalfont St. Giles: British Society for the History of Science, 1982. John Stuart Mill. “The Spirit of the Age” [1831]. Collected Works of John Stuart Mill. Ed. John M. Robson et al. Toronto: University of Toronto Press, 1981-91. Vol. 32, 227-316; Jacob Bronowski. Science and Human Values [1956]. New York: Harper & Row, 1975. William Eamon. Science and the Secrets of Nature. Books of Secrets in Medieval and Early Modern Culture. Princeton: Princeton University Press, 1994.
Scientific Coordination as Ethos and Epistemology
299
epistolary networks of learned humanists only gradually came to accommodate the exchange of scientific information and specimens;10 the international scientific congresses of the latter half of the nineteenth and early twentieth centuries were patterned on but not identical to the international diplomatic conferences of the late eighteenth and early nineteenth centuries.11 This essay explores how historical specificity and epistemological mission intersected in the international scientific collaborations of the late nineteenth century, including the Carte du Ciel, the Internationale Gradmessung, and the Transits of Venus expeditions, concluding with the reflections of philosopher-scientist Charles Sanders Peirce on the intertwined epistemology and morality of scientific coordination. We are less interested in providing a detailed account of any one of these collaborations, which have been well-treated elsewhere,12 than in using them together to exemplify a late nineteenth-century solution to the problem of constituting a scientific collectivity, a many-headed knower. In particular, we are interested in the ways in which the epistemological goals of this kind of collectivity demanded an ethos on the part of its members. It was not enough to subscribe officially to the objectives and procedures of an international collaboration; values had to be internalized and sometimes sacrifices made when these values – which were at once moral and epistemological – conflicted with those that informed research within a different scientific collectivity. It is part of the historical location of the international collaboration as one kind of scientific collectivity that it did not always dovetail smoothly with other kinds – no more than the values and mores of different subcultures always harmonize with one another. Prima facie evidence that the values at 10
11
12
Paula Findlen. Possessing Nature. Museums, Collecting, and Scientific Culture in Early Modern Italy. Berkeley and Los Angeles: University of California Press, 1994; Justin Stagl. A History of Curiosity. The Theory of Travel, 1550-1800. Chur: Harwood, 1995. Éric Brian. “Transactions statistiques au XIXe siècle. Mouvements internationaux de capitaux symboliques.” Actes de la recherche en sciences sociales 145 (décembre 2002): 34-46; Aant Elzinga and Catharina Landström, eds. Internationalism in Science. London: Taylor Graham, 1996. Charlotte Bigg. “Photography and Labour History of Astrometry. The Carte du Ciel.” The Role of Visual Representations in Astronomy. History and Research Practice. Ed. Klaus Hentschel. Thun and Frankfurt a.M.: Deutsch, 2000. 90-106; Ulrich Völter. Geschichte und Bedeutung der internationalen Erdmessung. Munich: Verlag der Bayerischen Akademie der Wissenschaften, 1963; John Lankford. “Photography and the Nineteenth-Century Transits of Venus.” Technology and Culture 28 (1987): 648-57; Jimena Canales. “Photogenic Venus. The ‘Cinematographic Turn’ and Its Alternatives in Nineteenth-Century France.” Isis 93 (2002): 585-613.
300
Peter Galison/Lorraine Daston
stake in these clashes were at least in part moral is that breaches occasioned not only disappointment but also indignation, a sign that what was at stake was more than a diplomatic compromise among delegates of individual nation states. The morality in question derived from the epistemological challenge such sustained international collaborations were meant to meet: how to recognize and map phenomena that dwarfed merely human scales of time and space. The shape and path of a storm, the contour of an isotherm, the physiognomy of a landscape, the geographic distribution of a species – these were phenomena, all new objects of inquiry in the nineteenth century, whose regularities or very existence was hidden to the observer located too firmly in the here and now, whose field of vision was circumscribed by the humanly meaningful co-ordinates of a locale or lifetime.13 Even those phenomena that had been studied earlier by international cooperations, such as the transits of Venus,14 were investigated differently when made the objects of sustained rather than occasional collaborations that strove to make protocols, instruments, and observers commensurable with one another. Improved means of communication and transportation were the preconditions for but not the causes of these ever wider and more densely woven networks of observers. As the several abortive eighteenth-century attempts to build observer networks in meteorology bear witness, it was not enough to recruit participants or even distribute standardized instruments; the observers themselves must be socialized and standardized in order for the networks to cohere and endure.15 This is the juncture at which ethos met epistemology. 13
14 15
Susan Faye Cannon. “Humboldtian Science.” Science in Culture. The Early Victorian Period. New York: Science History Publications, 1978. 73-110; Michael Dettelbach. “Global Physics and Aesthetic Empire. Humboldt’s Physical Portrait of the Tropics.” Visions of Empire. Voyages, Botany, and Representations of Nature. Ed. Peter H. Reill and David Philip Miller. Cambridge: Cambridge University Press, 1996. 258-92; Marie-Noëlle Bourguet. “Landscape with Numbers. Natural History, Travel and Instruments in the Late Eighteenth and Early Nineteenth Centuries.” Instruments, Travel and Science. Itineraries of Precision from the Seventeenth to the Twentieth Century. Ed. idem, Christian Licoppe, and H. Otto Sibum. London and New York: Routledge, 2002. 96-125. Harry Woolf. The Transits of Venus. A Study in Eighteenth-Century Science. Princeton: Princeton University Press, 1959. Gustav Hellmann. “Die Entwicklung der meteorologischen Beobachtungen in Deutschland von den ersten Anfängen bis zur Einrichtung staatlicher Beobachtungsnetze.” Abhandlungen der Preussischen Akademie der Wissenschaften, Physikalisch-Mathematische Klasse 1 (1926): 1-25; Robert Marc Friedman. Appropriating the Weather. Vilhelm Bjerknes and the Construction of a Modern Meteorol-
Scientific Coordination as Ethos and Epistemology
301
It was also the birthplace of new scientific objects. Many-headed knowers not only study nature in ways different than solitary knowers do; they also come to know things that located individuals, no matter how brilliant and diligent, cannot. The parallax of Venus during a transit, the exact shape of the earth, the distribution of stars in the sky and the slow changes in this configuration over millennia – these are phenomena that demand not only collaboration but exquisite coordination among scattered observers. Units of measure, instruments, clocks, as well as reaction times, note-taking habits, perceptual judgments, and certain values must all be synchronized. Otherwise the parts of the puzzle will prove incommensurable; the new scientific object will remain invisible. This is not a claim about metaphysical realism or its antonyms: we take for granted that storm systems and the shape of the earth exist whether or not they are successfully observed. Rather than beating the dead horse of metaphysical realism, we are concerned here with the intrinsically collective conditions of knowability. There are several kinds of such epistemological collectives, each specific to its age. The late nineteenth century offered peculiarly propitious conditions for a certain variety of planet-spanning scientific collaboration: colonial and commercial enterprises cast an ever more thickly woven net of connections over ever more of the globe. The first successful Atlantic telegraph cables began operation in 1866;16 the General (later Universal) Postal Union was created in 1874;17 a flurry of international meetings (themselves made possible by faster and more reliable communication and transportation) met to standardize units of measurement for science and commerce.18 Equally important were the administrative
16 17 18
ogy. Ithaca: Cornell University Press, 1989; James Rodger Fleming. Meteorology in America, 1800-1870. Baltimore and London: Johns Hopkins University Press, 1990. Tom Standage. The Victorian Internet. The Remarkable Story of the Telegraph and the Nineteenth-Century’s On-Line Pioneers. New York: Walker, 1998. James D. Cotreau. The Historical Development of the Universal Postal Union and the Question of Membership. Boston: n. publ., 1975. Simon Schaffer. “Late Victorian Metrology and Its Instrumentation. A Manufactory of Ohms.” Invisible Connections. Instruments, Institutions, and Science. Ed. Robert Bud and Susan E. Cozzens. Washington: Spie Optical Engineering Press, 1992. 23-56; idem. “Metrology, Metrification and Victorian Values.” Victorian Science in Context. Ed. Bernard Lightman. Chicago and London: University of Chicago Press, 1997. 438-74; Christophe Bonneuil. “The Manufacture of Species. Kew Gardens, the Empire, and the Standardisation of Taxonomic Practices in Late Nineteenth-Century Botany.” Instruments, Travel and Science. 189-215; Peter Galison. Einstein’s Clocks, Poincaré’s Maps. Empires of Time. New York: W.W. Norton, 2003.
302
Peter Galison/Lorraine Daston
skills of coordination acquired by expanding empires, orchestrating the activities of far-flung envoys and officials from a remote capital. Industrial materials and methods provided a model for the efficient organization of large laboratories and research expeditions. These technologies of steam, wire, and paper did not cause scientific collaborations, nor even make them possible. Long before the telegraph and the steamship, savants had been in communication with one another, by letter and in person. But the expansive ambitions of late nineteenth-century empire and commerce did enable and inspire scientists to think on a different scale, both geographical and chronological. European colonial outposts in the southern hemisphere made mapping, terrestrial and celestial, more comprehensive and sustained than eighteenth-century scientific travelers could ever have imagined. Diplomatic experience in negotiating international treaties was transferred, along with the rhetoric and reality of conflicting national interests, to international scientific congresses – themselves an innovation in this period. The resources of colonial administrations were put at the disposal of traveling scientists. Global reach recreated international science in its own image and enlisted it to serve its own commercial, military, and ideological purposes. Our examples of this new species of scientific collectivity and its ethical and epistemological implications are, in order of exposition, the Carte du Ciel, the Internationale Gradmessung, and the Transits of Venus (the polyglot titles of the projects already signal their international scope). This list is meant to be exemplary, not exhaustive; many more such late nineteenth-century scientific collaborations might be added. We have chosen these three to make a point about what many-headed knowers can know and how. The Carte du Ciel (begun in 1892) aimed to map the entire heavens, presenting earth-bound astronomers with their first complete picture of the distribution of all heavenly bodies down to the fourteenth magnitude (with a catalogue of stars down to the eleventh magnitude), and to supply future astronomers thousands of years hence with the means for detecting changes in the heavens since ca. 1900. Both spatial distribution and temporal development were new objects of scientific inquiry, which depended essentially on a scientific collectivity extended across continents and centuries. The ambitions of the Internationale Gradmessung (begun in 1886) were purely synchronic, but also vast: the determination of the shape of the earth by measuring variations in the gravitational constant all over its surface. These projects delineated objects of inquiry that previously could have been vaguely circumscribed (“all the stars in the sky”), but that had not yet been made scientifically knowable. In contrast, the Transits of Venus
Scientific Coordination as Ethos and Epistemology
303
expeditions (1874, 1882) sought to specify an entity already well defined since Copernicus: the distance from the earth to the sun. Known as the astronomical unit, it is the basis for establishing all absolute distances within the solar system. Yet as in the case of the Carte du Ciel and the Internationale Gradmessung, only the most meticulous coordination could make this object of inquiry precisely measurable. In the context of large-scale scientific collaborations, both literal and figurative standpoints are at issue. The literal standpoint of an observer in space and time can make certain entities invisible: storm systems, isotherms, the emergence and disappearance of stars, the distribution of an organic species. Hence networks of dispersed observers are formed to track these scientific objects that transcend human scales of the here and now. Many nineteenth-century scientific collaborations in meteorology, geodesy, statistics, biology, and astronomy aimed at tracking these global or cosmic objects by uniting the efforts of local observers into a kind of Leviathan or super-observer. This is where the figurative sense of standpoint enters. In order to co-ordinate the local observers into a global network, other kinds of peculiarities must be planed away. These might inhere in a group – e.g. national traditions of instrumentation and data reduction techniques – or in the individual – e.g. personal equation, greater or lesser zeal for the project in question. All of these forms of being situated could and did interfere with the communication and commensurability of observations. International commissions formed to steer the Carte du Ciel, the Gradmessung, and other scientific collaborations wrangled at length over how best to erase the influence of these figurative standpoints, in the service of overcoming literal ones. Only the many-headed knowers created by painstaking coordination – not erasing – of perspectives could detect scientific objects on a planetary or even cosmic scale. Mapping the Heavens From 16-25 April 1887 fifty-eight astronomers from sixteen countries plus three colonies met in Paris at the invitation of Admiral E.B. Mouchez, director of the Paris Observatory and the Paris Academy of Sciences, to plan what one contemporary called “the greatest venture yet undertaken in astronomy,”19 namely a complete photographic map of the 19
Julius Scheiner. Die Photographie der Gestirne. Leipzig: Wilhelm Engelmann, 1897. 311.
304
Peter Galison/Lorraine Daston
sky, including all stars to the fourteenth magnitude.20 Only the combined and prolonged efforts of almost a score of observatories in both the northern and southern hemispheres could produce what promoters hailed as an “imperishable monument,” a photographic record of “the authentic state of the universe visible from the earth at the close of the nineteenth century.”21 The proportions of the project were indeed monumental in every sense: eighteen observatories around the world, from Helsingfors at +60.9 degrees latitude to Melbourne at -37.522 labored for decades – publication of the catalogue was not completed until 196423 – to amass charts projected in 1912 to stack 32 feet high and weigh about 4,000 lbs.24 Armed with this snapshot of the sky circa 1900, future astronomers would be able, it was hoped, to detect changes in the heavens which unfolded on too long a time scale to be perceptible within a short human lifetime: appearance of new stars, nebulae, and comets, the telltale motion of as yet undiscovered planets, the extended periods of variable stars, the incremental proper motions of the so-called fixed stars.25 By 20
21 22
23
24 25
A star catalogue down to the eleventh magnitude was also planned as part of the Carte du Ciel project. On the history of nineteenth-century astrophotography in general, cf. Daniel Norman. “The Development of Astrophotography.” Osiris 5 (1938): 560-94; Dorrit Hoffleit. Some Firsts in Astronomical Photography. Cambridge: Harvard College Observatory, 1950; John Lankford. “The Impact of Photography on Astronomy.” Astrophysics and Twentieth-Century Astronomy to 1950. Ed. Owen Gingerich. Cambridge and New York: Cambridge University Press, 1984. 16-39. On the Carte du Ciel project in particular, cf. Institut de FranceAcadémie des Sciences. Congrès astrophotographique international tenu à l'Observatoire de Paris pour le levé de la Carte du Ciel. Paris: Gauthier-Villars, 1887; Albert G. Winterhalter. The International Astrophotographical Congress and A Visit to Certain European Observatories and other Institutions. Report to the Superintendent. Washington: Government Printing Office, 1889; Ernest B. Mouchez. La Photographie astronomique à l'Observatoire de Paris et la Carte du Ciel. Paris: Gauthier-Villars, 1887; Herbert Hall Turner. The Great Star Map. New York: E.P. Dutton, 1912; Suzanne Débarbat et al., eds. Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Dordrecht, Boston, and London: Kluwer, 1988. Camille Flammarion. “La photographie céleste à l'Observatoire de Paris.” Revue d'Astronomie Populaire 5 (1886): 55. For the zone assignments of individual observatories, cf. Nathy P. O’Hora. “Astrographic Catalogues of British Observatories.” Mapping the Sky. 136. For a tabulation of the final contributions, cf. Lankford. “Impact.” 30. Commission 23 of the International Astronomical Union, established in 1919 to oversee the Carte du Ciel, was dissolved in 1970. Théo Weimer. “Naissance et développement de la Carte du Ciel.” Mapping the Sky. 30. Turner. Great Star Map. 145. The project ultimately produced some 22,000 plates. Lankford. “Impact.” 30. On the expected advantages of the map, cf. Mouchez. Photographie. Chapters 3-4.
Scientific Coordination as Ethos and Epistemology
305
uniting astronomers around the world and across generations, the Carte du Ciel aspired to nature’s own Brobdingnagian scale. Although photography made this immense project conceivable, and although its supporters at times invoked the ideals of mechanical objectivity,26 it was the ethos of the map constructed by many hands that guided the makers of the Carte du Ciel. Astrophotography promised speed, permanence, and authenticity, but not globality and uniformity. As the deliberations of the 1887 International Congress and of subsequent meetings (1889, 1891, 1896, 1900, 1909) of the Permanent Committee make clear,27 the intricate co-ordination of telescopes, photographic plates, micrometric measurements, and myriad other details to ensure that the parts of the map would be commensurable required that participants relinquish control not only over instruments and methods, but also over the choice of research area for years to come. The debate over the kind of telescope to be used in the photographic work of the Carte du Ciel shows how dearly uniformity and globality were sometimes purchased. Although British astronomers such as Andrew Ainslie Common and Isaac Roberts had pioneered stellar photography using reflecting telescopes,28 and although, unlike refractors, reflectors could be used for both visual and photographic observations, it was Common and Roberts who recommended that “the reflector should yield to the refractor in a work to be undertaken in concert.” Common defended the reflector as “the best instrument in all respects for celestial photography,” but its proper use required “long and careful experiments,” in contrast to the refractor, “whose manipulation was easily learned.” Roberts seconded the point that the reflector “required the exercise of great care and patience, and a thorough personal interest on the part of the observer using it. In the hands of such a person it yielded 26
27
28
On mechanical objectivity, cf. Lorraine Daston and Peter Galison. “The Image of Objectivity.” Representations 40 (1992): 81-128. There were references to a map made “by photography alone and without the intervention of any human errors” and to replacing “the personality [i.e., the personal equation] of the observer by the impersonality of the plates.” Camille Flammarion. “Le Congrès astronomique pour la photographie du ciel.” Astronomie 6 (1887): 161-69, on 163; T.N. Thiele quoted in Winterhalter. The International Astrophotographical Congress. 59. Cf. Winterhalter. The International Astrophotographical Congress; Institut de France-Académie des Sciences. Congrès; also the irregularly published Bulletin du Comité Permanent International pour l'Exécution Photographique de la Carte du Ciel. On Common’s and Roberts’s photographic work with reflectors, cf. John Lankford. “Amateurs and Astrophysics. A Neglected Aspect in the Development of a Scientific Specialty.” Social Studies of Science 11 (1981): 275-303.
306
Peter Galison/Lorraine Daston
excellent results, but in other hands it might be a bad instrument.” Neither Common nor Roberts was confident that all observers in so large and international an undertaking would command the necessary skill and patience; therefore, “for the sake of securing uniformity in the operations of a large number of astronomers,” they urged the adoption of a refractor like the one used by the brothers Prosper and Paul Henry at the Paris Observatory.29 What was being sacrificed here was not simply the expense of building a new telescope to the specifications laid down by the international congress,30 nor British pride in their illustrious tradition of reflectors, the telescope invented by Newton. At stake was also skill, and the accuracy vouchsafed by skill, for skill in the operation of reflectors was too local a peculiarity to be safely standardized. Where nearly a score of observatories and hundreds of observers had to mesh their methods and results into a seamless whole, there was no room for any idiosyncrasy, including the idiosyncrasy of superior accuracy.31 The flattening pressure of uniformity was also exerted on superior precision in the measurement of the plates to determine stellar positions. By resorting to the labor of poorly paid schoolboys and legions of volunteers, Oxford astronomer H.H. Turner was able to complete at least that part of the astrographic catalogue assigned by the International Congress to his observatory by 1911, although the Oxford contribution to the photographic map was never completed.32 Impatient with those observatories still in arrears in delivering their portion of the Carte du Ciel, Turner chided his colleagues elsewhere for indulging in 29 30
31
32
Winterhalter. The International Astrophotographical Congress. 18. The price of each telescope was estimated at FF 40,000, plus an additional FF 21,000 for the dome and measuring apparatus. Flammarion. “Le Congrès astronomique.” 167-68. Most of the refractors used in the Carte du Ciel were made either by Grubb in Dublin (seven) or Gautier in Paris (nine). O’Hara. “Catalogues.” 136; Patrick A. Wayman. “The Grubb Astrographic Telescopes.” Mapping the Sky. 139-42. Throughout the 1887 Congress there were tensions between those who demanded “perfect identity” and those who defended superior instruments or methods. For example, Folie, director of the Brussels Observatory, and Janssen, President of the Paris Academy of Sciences, pleaded that more powerful but non-standard instruments like the Meudon telescope be permitted to take part in the mapping. Janssen also noted that it would violate long-standing practice to dictate standards to the best instrument-makers in the various countries participating in the Carte du Ciel: “lorsque l'on s'adresse à des artistes de talent, on leur laisse habituellement une grande latitude pour les détails de la construction.” Institut de France-Académie des Sciences. Congrès. 26, 46. O’Hora. “Catalogues.” 137.
Scientific Coordination as Ethos and Epistemology
307
what he deemed an excess of precision in the measurement of their plates. The whole Carte du Ciel was in jeopardy as a result; work that under the best of conditions would have taken twenty years now threatened to stretch out for at least forty: The fact is that the necessity for strenuous economy has not been sufficiently realised: some of the larger observatories strained at an accuracy scarcely possible even for them; and their weaker brethren, in attempting to copy their example, have been left far behind. Moderation and self-denial are just as necessary in astronomical work as in other walks of life.33
Turner’s stern appeal to “self-denial” in the context of getting on with the Carte du Ciel project was of a qualitatively different kind than the self-restraint preached by the advocates of mechanical objectivity, albeit no less moralized in tone. Individual scientists ought to exercise self-restraint in judgment and interpretation, lest their voices drown out nature’s own.34 In contrast, individuals or, more often, research groups were admonished to practice self-denial in their choice of equipment and methods, lest local peculiarities jeopardize the joint effort to grasp nature as a whole. The levels of sacrifice demanded by the scientific collectivity were several: the cost in time and money of new instrumentation and training, the relinquishing not only of tried and true but also sometimes of more accurate methods, the substitution of efficiency for painstaking precision, the monopolization of resources and personnel for long periods of routinized labor, the steadfast resistance to the temptation to neglect old collaborative commitments in pursuit of an exciting new discovery. These sacrifices were particularly grave for the smaller observatories participating in the Carte du Ciel project: for example, the Australian observatories of Sydney, Perth, Melbourne, and Adelaide took 80 years to complete their three assigned zones of the sky (18% of the entire sky), at the price of limiting other investigations, particularly in the enormously fruitful fields of astrophysics and spectroscopy.35 Some astronomers found the sacrifices required by the Carte du Ciel too great. T.N. Thiele, director of the Copenhagen Observatory, declined to take responsibility for a zone, for he foresaw that it would conscript the observatory into years of monotonous labor and jeopard33 34 35
Turner. Great Star Map. 75. Claude Bernard. Introduction to Experimental Medicine [1865]. Trans. Henry Copley Greene. New York: Dover, 1957. 22f. Graeme L. White. “The Carte du Ciel – The Australian Connection.” Mapping the Sky. 48; cf. Lankford. “Impact.” 32, on the converse advantages to American observatories which did not participate in the Carte du Ciel.
308
Peter Galison/Lorraine Daston
ize the precision measurements that were his chief interest.36 But even those who refused to compromise their autonomy by participating in the Carte du Ciel acknowledged the moral authority of the collaboration. Edward Pickering, director of the Harvard Observatory, preferred to conduct his own astrophotographic survey when the 1887 International Congress did not adopt his recommendations concerning instruments and techniques.37 Yet Pickering felt obliged to defend his defection to Admiral Mouchez in 1889, opposing scientific individualism and diversity to the Congress’s call for collaboration and standardization: I desired to avoid any appearance of dissatisfaction with the decisions of the International Congress, as I was, as I still am, unwilling to criticize the plans, prepared with such care by so large a number of representative astronomers. But I cannot think it best for the promotion of science to abstain on this account from the attempt to carry into execution a series of observations which has seemed desirable to me for many years . . . In the present state of astronomical photography, it seems to me there is no danger of any serious waste of labor in the repetition of photography of the same regions taken by different methods, each of which may prove to have special merits.38
Given that at this time astrophotographic techniques were being refined yearly, Pickering’s plea for flexibility and experimentation was reasonable from the standpoint of securing the most accurate star charts possible. Yet for those astronomers who had volunteered the rest of their working lives to the Carte du Ciel, his breezy dismissal of commensurability must have seemed the height of egotism. Coordinating Coordinates Mapping the skies was a suitably ethereal undertaking; the determination of the earth – its dimensions and shape – however followed at least in part altogether practical demands. By the early 1860s, ocean navigation was in full-ahead expansion, the rigorous establishment of political boundaries was increasingly vexed, and re-worked maps for Europe and beyond were a major concern. Railroads and miners wanted excellent maps; so too did the colonizers as they began carving up Africa, 36 37
38
Leif Kahl Kristensen. “T.N. Thiele and the Carte du Ciel.” Mapping the Sky. 59-63. Pickering did however serve on the photometric commission of the Permanent International Committee of the Carte du Ciel. Lankford. “Impact.” 38. For the full text of Pickering’s recommendations to the International Congress of 1887, cf. Winterhalter. The International Astrophotographical Congress. 55-58. Letter of Pickering to Mouchez, 14 August 1889. Pickering Papers, Harvard University Archives, UAV 630.14, ser. A-9, 3f.
Scientific Coordination as Ethos and Epistemology
309
South America, and Southeast Asia. Beyond the practical, measuring the figure of the earth mattered for an understanding of the law of gravity and the basic constants of solar system distances. Peirce, working for the Coast and Geodetic Survey, insisted on the point: Although in laying out the plan of a geodetical survey the relative utility of the knowledge of different quantities ought to be taken into account, and such account must be favorable to pendulum work, yet it is also true that nothing appertaining to such a survey ought to be neglected, and that too great a stress ought not to be put on the demands of the practically useful. The knowledge of the force of gravity is not a mere matter of utility alone, it is also one of the fundamental kinds of quantity which it is the business of a geodetical survey to measure.39
Peirce was by no means alone in his attention to the abstract as well as the concrete. For centuries leading lights of natural philosophy had been concerned with the shape of the earth. Newton had famously used the measurement between Paris and Amiens in order to determine the shape of the earth - the earth, he argued, should be flatter near the poles because of the earth’s spin on its axis. Despite all this interest, as local or national measurements proceeded in the nineteenth century, it had become ever clearer that there were discrepancies and gaps between these partial efforts. For example, various longitude surveys concluded that the earth was more or less flattened at the poles. Astronomical measurements clashed with geodetic ones – in one case they clashed over the longitudinal difference between Milan and Turin by 31.29”. And when different observers swung pendula to determine the strength of gravity (a quantity that varied with latitude because of the spinning of the earth) their results differed by as much as 3%. For all these reasons – as well as for the more abstract mathematical-physical desire to know the detailed shape of the earth – scientists from many countries had made measurement of the earth a matter of international concern.40 Among other advocates, C.F. Gauss had hoped for a vast chain of interlocking trigonometric measurements to map the world. Johann Jacob Baeyer had begun this task first for the Prussian military, then as part of a European Gradmessung, and with no small amount of difficulty helped form a General Conference that, from 1864 through 1912, met some seventeen times.41 This coordination would not be trivial, as Baeyer’s “General Report” made clear already in 1862: 39
40 41
Charles S. Peirce. “Six Reasons for the Prosecution of Pendulum Experiments” [1882]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 359f. Völter. Geschichte. 7-9. Ibid. 7-9.
310
Peter Galison/Lorraine Daston
Since the greatest possible similarity in form must be desired, the question arises as to whether, in order to achieve this goal, it would not already be best to design a general working plan and to agree upon this at a general conference. As advisable as this may seem at first glance, on closer inspection one comes up against difficulties which, if agreed upon too soon, in all probability could not be cleared up. For the state of measurement in the various countries is no less variable than the available means and personnel, so that one is indeed everywhere obliged to adjust to particular relations and conditions, of which one cannot assume that they may be treated in the same way.42
There were difficulties coordinating procedures, standards, and equipment. Even the individual observers found it hard to execute the work because of an amalgam of physiological, meteorological, and instrumental difficulties. In 1865 Baeyer reported that their original intention to conduct night observations had been thwarted first because his eyes were tired after working all day and they needed rest in order to be fresh for the next set of measures. Second, the lighting equipment for the cross hairs differentially warmed the parts of the instrument leading to errors.43 Problems multiplied when inter- as well as intra-individual observations had to be coordinated, especially among various national scientific équipes. Thirteen years later, in 1875, friction among the participating nationalities had only increased. At a meeting called to standardize the form for recording data, tension surfaced the moment these differences became apparent. French-German relations were already severely strained by the Franco-Prussian war in which the Prussians had soundly pummeled the French. Peirce was present for the assembly, so too were many of the great astronomers of the day, including Swiss astronomer Adolph Hirsch and his French colleague Hervé Faye. The Frenchman Yvon 42
43
“Da hierbei eine grösstmögliche Gleichformigkeit wünschenswerth sein muss, so entsteht die Frage, ob nicht, zur Erzielung derselben, schon jetzt ein allgemeiner Arbeitsplan zu entwerfen und auf einer allgemeinen Conferenz zu vereinbaren wäre. Wie zweckmässig dies auch auf den ersten Blick erscheinen mag, so stösst man doch bei näherem Eingehen auf die Sache, auf Schwierigkeiten, welche bei einer zu frühzeitigen Vereinbarung aller Wahrscheinlichkeit nach nicht aus dem Wege geräumt werden können. Denn der Stand der Vermessungen ist in den verschiedenen Ländern nicht minder verschieden als die disponiblen Mittel und Kräfte, so dass man genöthigt sein wird sich thatsächlich überall nach besonderen Verhältnissen und Umständen zu richten, von denen man nicht annehmen kann, dass sie sich gleichartig behandeln lassen.” General-Bericht über die mitteleuropäische Gradmessung für das Jahr 1862. Berlin: Reimer, 1863. 5. J.J. Baeyer. “Bericht über den allgemeinen Standpunkt der Preussischen Vermessungen in Bezug auf die mitteleuropäische Gradmessung, und im Besonderen über die im Jahre 1864 ausgeführten Arbeiten.” General-Bericht über die mitteleuropäische Gradmessung für das Jahr 1864. Berlin: Reimer, 1865. 33
Scientific Coordination as Ethos and Epistemology
311
Villarceau launched one donnybrook when he dismissed the recording of probabilistic errors as “useless” in the registration of astronomical coordinates. At this, Hirsch reacted with hard-edged diplomacy: of course the French were not obliged to fill out every column of the agreed-upon forms. The Commission obliged no one to do so. But if the French did not calculate the probable errors – which Hirsch himself judged to be a very regrettable state of affairs – one would not be able to say what would happen in other countries.44 Their probabilistic honor besmirched, Astronomer Faye responded indignantly, protesting that not only had some of the most beautiful work on the subject been done by French mathematicians, but that the French astronomers prized the work highly indeed, and taught probability in their lectures. Counter-protested Hirsch: no possible offense was meant to M. Faye; Hirsch’s concerns touched only M. Villarceau. Hirsch at no time would have dreamt of sullying all French scholars. Perish the thought. The idea that his illustrious colleagues would have neglected probability calculus or the theory of least squares would not ever have crossed his mind.45 Sometimes the issue was not so much the universality of skill, but the validity of a particular kind of instrument. When one 1875 delegate enthusiastically backed the accuracy of a slightly modified Repsold Pendulum for the determination of mass (“it meets all the scientific demands”), Peirce rose to protest. He had discovered that the swinging of the pendulum gave rise to a distortion of the three-legged stand, so one would find a falsely shortened length of the pendulum.46 But whether one used the Repsold Pendulum or one of its competitors, swinging pendula, measuring time intervals, setting up stations at farflung sites in exceedingly demanding conditions was difficult work. The temptation to use automatic, self-recording devices was great – there were even some moderately successful trial attempts. After all, around the mid-1870s self-recording instruments were becoming more and more common in domains from medicine to long-distance timing signals for longitude determination. But the delegates were not persuaded. Automatic registration promised to eliminate – at least in principle – the variation among observers. Peirce, Hirsch, and their colleagues remained attached to the observer-run measurement.47 44 45 46 47
Verhandlungen der vom 20. bis 29. September 1875 in Paris vereinigten Permanenten Commission der Europäische Gradmessung. Berlin: Reimer, 1875. 56f. Ibid. Ibid. 59. Ibid. 60.
312
Peter Galison/Lorraine Daston
While they agreed to abandon automatic registration devices, the delegates unanimously concluded that for the Gradmessung to succeed they would need to provide an international apparatus. It was not that one needed a new, more accurate device. Those in current use were, in and of themselves, accurate to a very high degree. No, the problem was not that the current instruments were, taken separately, inaccurate. Instead, this international device would assist at the boundaries between the already-existing national networks of geodetic triangulation measurements – where it had become all-too noticeable that agreement was very difficult. This had sadly been the case when the Italian and Austrian officers met to establish a baseline at Udine. Furthermore, an international apparatus would offer an unambiguous choice for geodecists from the smaller European countries. At long last it would be possible to bring the European Net into a unified form, and to have all lengths expressed in identical units. If this were admitted, some of the leading figures argued, then the natural consequence would be the establishment of an international metrical standard mass. That mass should be located in the International Bureau of Weights and Measures, under the authority of observers from the various countries who were setting the base lines for their measurements.48 As Peirce put it in 1879, “The value of gravity-determinations depends on them being bound together, each with all the others which have been made anywhere on earth.”49 Many measurements demanded great skill, skills of instrument making, instrument maintenance, calculation, and data reduction. In the Carte du Ciel it is clear that the British astronomers could have proceeded with their reflectors, but given many other countries’ inability to build and maintain these finicky devices, the effort as a whole was quite likely to fail. The British ceded – as we have seen – and the refractor became the standard instrument. Villarceau may have exaggerated the difficulty of maintaining error bars, but his general point was quite clear: demand too much skill and you lose commensurability. Some five years after he had protested at the Germans’ insistence on the recording of errors, he came back with a similar protest about demands that seemed to exceed the abilities of the participants: Mr. Yvon Villarceau considers it important to differentiate the relative gravity determinations from the absolute ones, the commission having concerned themselves only with the latter. According to him, the absolute determinations present
48 49
Ibid. 62f. Charles S. Peirce. “Measurements of Gravity at Initial Stations in America and Europe” [1879]. Writings. Vol. 4, 81.
Scientific Coordination as Ethos and Epistemology
313
so many difficulties that these should only be made in a certain number of observatories, where all the necessary tools are available.50
For once, les Anglo-Saxons and the French seemed to agree. Efficiency, and the accompanying pragmatic demands of keeping the results on track, were not always easily compatible with the ultimate in measurement precision. On 5 May 1882, Major (later Colonel) John Herschel of the Trigonometrical Survey of India wrote to Prof. J.E. Hilgard, Superintendent of U.S. Coast and Geodetic Survey. The subject was the conference on gravity observations held in Washington, D.C., but the issue more general: I hold it to be a very lamentable thing that men of zeal, eager to advance science, should continue to be misled by the old school of physics into launching upon the difficult and precarious entreprise of absolute determination of gravity, generally in ignorance of the real difficulties of the research, and always indifferent to the utility of such determination. The German school is responsible for this.51
Herschel then went on to speak of his own, British-supported latitude work in India. He insisted that the same arguments and motives . . . point out the urgent need for economy in every detail of installation and observation – in the choice of stations and the buildings to be occupied, in the distribution of time to be taken up by the observations, and by the calculations respectively, so as to get, in short, as many results of a sufficient degree of accuracy, and no more, as possible within the year.52
Peirce appreciated Herschel’s pragmatic stance. As far as Peirce was concerned, it was not only a matter of the raw difficulty of absolute determinations of gravity, but more generally, whether the Germans’ insistence on precise least-squares analysis captured the real physical difficulties in estimating errors. Protocol, Peirce was essentially arguing, did not exhaust error: The history of pendulum observations abounds with inexplicable contradictions and anomalies indicative of unknown causes of error; and hardly a single observer 50
51 52
“Herr Yvon Villarceau hält es für wichtig, die relativen Schwerbestimmungen von den absoluten zu unterscheiden, mit welchen letzteren allein die Commission sich beschäftigt hat. Seiner Meinung nach bieten die absoluten Bestimmungen so viel Schwierigkeiten dar, dass man dieselben nur in einer gewissen Anzahl Sternwarten anstellen sollte, wo man alle nöthigen Hülfsmittel besitzt.” Centralbureau der Europäischen Gradmessung. Verhandlungen der vom 13. bis 16. September 1880 zu München abgehaltenen sechsten allgemeinen Conferenz der Europäische Gradmessung. Berlin: Reimer, 1881. 27. Reply of Major J. Herschel to Prof. J.E. Hilgard, 5 May 1882. Peirce. Writings. Vol. 4, 353. Ibid. Vol. 4, 354f.
314
Peter Galison/Lorraine Daston
has ventured to estimate the probable error of his result. Practically, the question of precision is cut by a variety of circumstantial exigencies; and it would seem best to leave it at the discretion of the observer, or director of the work.53
Perhaps there is a kind of national poetic appropriateness to these various responses to error estimation: the Germans wanted a rigorous and universal protocol, the Anglo-Americans took the matter to be one of individual, pragmatic choice, and the French found the whole idea of error estimation to be useless. Instruments that in theory were supposed to work well were in practice found to bend and twist in data-wrecking ways. Accurate national surveys clashed on the boundaries, automatic instruments left the delegates fearful of unanticipated error, and the hope – not often fulfilled – was for a universally-shared apparatus that would finally put paid to local variation. For the shape of the earth, for the determination of gravity at different points, the world-surveying project required a planet-covering net of observers. All those small-scale, partial measurements of the eighteenth and early nineteenth century had, by their discrepancies and gaps, made their inadequacy all too clear. The conditions for a “collective observer” to come into existence took decades, dozens of conferences, new forms of instrumentation, data reduction, and world-covering expeditions. Tracking Venus Mapping the heavens and determining the shape of the earth depended crucially on the commensurability of the results submitted by dispersed observers, which in turn depended on intricate coordination of observers, instruments, measurements, and protocols. Without such coordination, costly and nerve-racking though it might be, there would be no way of piecing the parts of the puzzle into a unified scientific object. In the case of the nineteenth-century Transits of Venus expeditions, the problem of coordinating observers, even those with the same training using the same instruments in the same place, proved almost insoluble, despite heroic efforts. The Transits of Venus expeditions mounted by France, Great Britain, Germany, the United States, and several other countries in 1874 and 1882 were perhaps the single most elaborate and expensive scientific 53
Charles S. Peirce. “General Remarks upon Gravity Determinations, by John Herschel.” Writings. Vol. 4, 368.
Scientific Coordination as Ethos and Epistemology
315
undertaking in a century distinguished by ambitious global scientific projects like the Carte du Ciel and the Internationale Gradmessung. The transits were considered to be the most reliable means of determining solar parallax and therefore the astronomical unit, which was in turn the basis for all other calculations of distances within the solar system. The transits occur only once every 121 years, and then in pairs eight years apart. Although the 1639 transit had been observed at least in England by Jeremiah Horrocks and William Crabtree, the idea of using transit observations from remote points on the earth’s surface to ascertain solar parallax was first advanced by Edmond Halley after he had observed a transit of Mercury on the stormy island of St. Helen in 1677.54 Halley’s proposal was put into practice only in 1761 and 1769, when astronomers fanned out to Pondicherry and Siberia, Cape Town and Vera Cruz, Santo Domingo and Lapland to track the path of Venus across the sun’s surface. The very nature of the phenomenon required a dispersed network of observers, who traveled vast distances with delicate instruments to make observations that only made sense when coordinated with one another. Bedeviled by war, illness, bad weather, and, above all, inexperience in observing the phenomenon in question, the eighteenth-century transit expeditions had produced maddeningly inconsistent and uncertain results.55 As the transits of 1874 and 1882 approached, astronomers were determined to improve upon the disappointing measurements of their predecessors. They intended to take full advantage not only of new technologies such as photography and the telegraph, but also of imperial grids of transportation, communication, and military organization. Their expeditions, self-consciously described as the continuation of the heroic efforts of the eighteenth-century observers, would extend the observing network commenced over a century earlier in time as well as space. Three methods were proposed to measure solar parallax during the transit, each dependent on a different, but equally finicky technology. The method of duration (sometimes called Halley’s method) measured the duration of the transit as observed from two widely separated stations; here exquisitely calibrated chronometers and equally well-calibrated observers were required to ascertain exactly when contact between Venus and the sun occurred. The heliometric method adapted an instrument originally used to measure the diameter of the sun’s disc to find the exact distance from the edge of Venus to the edge of the sun. 54 55
David Sellers. The Transit of Venus. The Quest to Find the True Distance to the Sun. Leeds: Maga Velda Press, 2001. 75-90, 104-18. Harry Woolf. The Transits of Venus.
316
Peter Galison/Lorraine Daston
Again, considerable skill was required to get good results: “It is a very troublesome instrument to manipulate, and the corrections due to the influence of temperature are extremely difficult to apply. Yet with great care there is little doubt that very accurate measurements can be made.”56 But many astronomers pinned their hopes on the third method, which took a sequence of photographs at intervals as brief as a second of the planet’s movement across the sun’s disc, using the “photographic revolver” proposed by French scientist Jules Janssen.57 Janssen’s colleague Faye praised the new photographic methods to the Paris Academy of Sciences: “This type of observation removes the observer, and with him the anxiety, the fatigue, the bedazzlement, the haste, the errors of our senses, in a word, the intervention – always suspect – of our nervous system.”58 The eye and brain of the observer would be replaced by a photographic plate and an electric telegraph; with such an apparatus in place, Faye boasted, even a child could outdo the most experienced astronomer.59 British, German, and American astronomers were perhaps less enthusiastic than the French about photographic methods, but almost all of the 1874 transit expeditions included some apparatus and personnel for making photographs of the transits: “This [photographic] method is looked forward to with much interest, because it is the first time that photography has been extensively employed in delicate astronomical measurements.”60 These great expectations were bitterly disappointed in the event. In his report on the results of the 1874 British transit expeditions, Astronomer Royal George Biddell Airy judged the photographic results not even worthy of analysis: The apparent uncertainty in the conclusions from the photographic registers has led extensively to the persuasion that it is unnecessary to record the photo56 57
58
59
60
George Forbes. The Transit of Venus. London and New York: Macmillan, 1874. 35. Jimena Canales. Sensational Differences. Individuality in Observation, Experimentation, and Representation (France 1853-1895). Ph.D. dissertation. Harvard University, 2003. Chapter 3. “Ce genre d’observation supprime l’observateur, et avec lui l’anxiété, la fatigue, l’éblouissement, la précipitation, les erreurs de nos sens, en un mot l’intervention toujours suspect de notre système nerveux.” Hervé Faye. “Sur l'observation photographique des passages de Vénus et sur un appareil de M. Laussedat.” Recueil des mémoires, rapports et documents relatifs à l'observation du passage de Vénus sur le soleil. Paris: Firmin Didot Frères, 1874. 178. Hervé Faye. “Rapport sur le rôle de la photographie dans l'observation du passage de Vénus.” Recueil des mémoires, rapports et documents relatifs à l'observation du passage de Vénus sur le soleil. Paris: Firmin Didot Frères, 1874. 228, 232. George Forbes. The Transit of Venus. 31.
Scientific Coordination as Ethos and Epistemology
317
graphic operations with the utmost detail . . . The ardour of the Observers has been much cooled by the apparent general failure of the photographic principle, and they were unwilling to spend further time on those reductions.61
The director of the Berlin Observatory Foerster estimated that the probable error of the distances between the centers of Venus and the sun measured from the photographic plates was about five times as great as that derived from the heliometric results. He traced the problem to the instantaneity of the photographs: The reason for this oscillation is that, in almost instantaneous photographic images produced, the current phase of atmospheric oscillation is photographed, while in micrometric measurement a good observer avoids them by fixing a median position for the images.62
Here the vaunted virtues of photography were turned into vices: automaticity and instantaneity recorded every minute fluctuation in the air, fluctuations that a seasoned observer knew how to smooth out. Even the French conceded defeat and supported the recommendation of the international conference, convened in Paris in a last-minute attempt to better coordinate the national expeditions of 1882, to minimize photographic efforts in the second, all-important attempt.63 This threw the weight of responsibility back upon the human, alltoo-human observers. Past experience, both of the eighteenth-century and 1874 expeditions, had shown how difficult the observations were to make under often makeshift conditions in remote locales with dodgy weather. The German mission sent to the Auckland Islands did not glimpse the sun for the entire month of November; on December 9th, “the, for us, so important day,” the rain abated only long enough to afford a brief glimpse of the transit.64 The French mission in Patagonia similarly complained of “the harshness of a detestable climate,” which 61
62
63 64
George Biddell Airy, ed. Account of Observations of the Transit of Venus, 1874, December 8, Made under the Authority of the British Government. And of the Reduction of the Observations. London: Her Majesty’s Stationery Office, 1881, Appendix V: “Photographic Observations of the Transit of Venus.” 14, 19. “La raison de cette inferiorité est que, dans des images photographiques presque instantanées, la phase actuelle des oscillations atmospheriques est photographiée, tandis que, dans les mesures micrométriques, un bon observateur s’en débarrasse en fixant pour les images une position moyenne.” Ministère de l'Instruction Publique et des Beaux-Arts. Conférence internationale du passage de Vénus. ProcèsVerbaux. Paris: Imprimerie Nationale, 1881. 6. Ministère de l’Instruction Publique et des Beaux-Arts. Conférence internationale du passage de Vénus. 24. Arthur Auwers, ed. Die Venus-Durchgänge 1874 und 1882. Bericht über die Deutschen Beobachtungen. Berlin, 1887-98. Vol. 1, 175.
318
Peter Galison/Lorraine Daston
impeded their work.65 The cooperation of the local authorities as well as that of the elements was a further precondition for success. The Mexican government had to lay a telegraph cable, build a road, and station a squadron of soldiers to enable the French mission to get on with their observations undisturbed;66 the British mission on the Hawaiian island of Waimea insured peace and quiet by posting sentries around the observing station and having “the grounds ‘tabooed,’ or, as they call it, ‘kapu,’ for the day.”67 Yet even when both weather and local inhabitants favored the observers, making the observations was often a vexed business. Captain G.L. Tupman, R.M.A., superintendent of the 1874 British expeditions and himself charged with the mission to Honolulu, reported optimal conditions on the day of the transit, 9 December 1874: the sky was cloudless. Her Majesty Queen Kapiolani of Hawaii had obligingly ordered her subjects (who had climbed trees and rooftops to observe the observers) “that ‘Silence’ must be maintained;” all the instruments were in readiness. Yet Tupman, bent over his telescope, was greatly distressed nonetheless to miss the all-important first contact of Venus with the sun’s surface: “After 20 seconds which I have recorded I am perfectly certain the contact was passed, established completely – not ‘contact’ properly speaking, for that implies some definite instant never observed.”68 The four points of contact between planet and sun’s perimeter – internal and external ingress, internal and external egress – so essential for the determination of parallax, were anything but sharply defined visual events. The French observers in Patagonia devised the following Cartesian criteria of contact: 1. The beginning of uncertainty (Le commencement du doute); 2. The presumed moment of contact (L’instant présumé du contact); 3. The end of uncertainty (La fin du doute).69
Their colleagues stationed on Chubut cautioned that the values they had obtained, albeit convergent, could not be treated as entirely accurate: “the observation is not in itself susceptible of great precision.”70
65 66 67 68 69 70
“la rudesse d’un climat détestable.”Académie des Sciences. Passage de Vénus du décembre 1882. Rapports préliminaires. Paris: Gauthier-Villars, 1883. 60. Ibid. 30. Airy, ed. Account of Observations of the Transit of Venus. 241. Airy, ed. Account of Observations of the Transit of Venus. 44. Académie des Sciences. Passage de Vénus. 65. “l’observation en elle-même n’est pas susceptible d’une grande précision.” Ibid. 81.
Scientific Coordination as Ethos and Epistemology
319
Almost all of the observers had been rigorously trained beforehand using artificial transit machines set up at observatories in Paris, Washington, Greenwich, and Berlin. Although the very nature of the transits of Venus, with luck visible only twice within a human lifetime, prevented the accumulation of observational experience by the usual means, it was hoped that practice with the models would hone the eyesight and judgment of the expedition members. “To make a really good observation of this contact,” intoned the official instructions issued to the American observers, two conditions are essentially necessary to all which have been described. The observer must have had some previous practice in observing first contacts, and must know exactly where to look for the contact. The first condition can be well fulfilled by the artificial transit of Venus apparatus, of which it is intended to have one or more available for observers.71
Astronomers were confident that the egregious errors of the 1761 and 1769 expeditions might, with sufficient advance training, be reduced from fifteen to two seconds.72 Yet when faced with the real transit, many well-schooled observers were dismayed to discover that what they saw did not resemble the familiar model. It was these divergences that persuaded Tupman he had missed first contact, despite straining every nerve to catch it: “Everything hitherto having so closely resembled the appearances in the model, I felt certain that I had missed the contact while focusing, although I could not understand how it could have occurred so much sooner than I expected.”73 Some of the German observers on the island off the coast of the Chinese port city Chih-fu were confused by the ring of light surrounding Venus caused by the illumination of its atmosphere as it passed in front of the sun: Because of these very effects, Reimann found the appearance so completely dissimilar from what he was familiar with from the model, that he was utterly unable to recognize the moment to which the observer’s attention was primarily directed . . . Quite the opposite, Valentiner was surprised by the complete similarity of the actual appearance to the representation by the model, and he found the main moment of observation to be precisely determined.74
71 72 73 74
U.S. Transit of Venus Commission. Instructions for Observing the Transit of Venus, December 6, 1882. Washington, D.C.: Government Printing Office, 1882. 40. George Forbes. The Transit of Venus. 54. Airy, ed. Account of Observations of the Transit of Venus. 43. “Reimann fand wegen der Mitwirkung desselben die Erscheinung so gänzlich unähnlich der ihm vom Modell her bekannten, dass er das Moment, auf welches die Aufmerksamkeit der Beobachter hauptsächlich hingelenkt war, gar nicht zu erkennen vermochte . . . Ganz entgegengesetzt wurde Valentiner durch vollkommene Ähn-
320
Peter Galison/Lorraine Daston
Even observers who had scrupulously practiced with the artificial transit machines found themselves confused by the actual phenomenon, even to the point of disagreement as to whether the observed appearances in situ resembled those of the model or not. All that could be done was to conserve attention, as an athlete husbands strength before a race: It is essential that the observer should allow his eye nearly perfect repose for several minutes before the contact. It is quite right and proper that he should take a general view of the phenomenon at short intervals, and note the appearance presented by the outline of the planet, but he should not exercise his eye and attention in endeavoring to make any difficult observation.75
Faced with these results, instructions to observers on the second, 1882 expeditions underscored the importance of writing down immediate, detailed, and inalterable descriptions, preferably with drawings, of exactly what each observer had seen. Probable errors might later be noted and corrected, but the original, unvarnished impressions and measurements must be preserved at all costs.76 These accounts, carefully kept and duly published in their entirety, testify to the bewildering diversity of perceptions among the observers, even among those who had been trained on the same models and who were observing side-by-side at the same station. Neither photography nor electric telegraphy nor artificial transit machines could dispense with the observer or eliminate troubling differences among observers. Artificial star machines, of which the transit models were the most important variety, had first been introduced in observatories in the 1850s and 1860s to measure and hopefully to eliminate the absolute personal equations of individual observers. Since Bessel’s analysis of the personal equations of himself and his colleagues Argelander, Struve, and Walbeck in 1820,77 several observatories had regularly computed the relative personal equations of pairs of observers. When in 1858 J. Hartmann invented an artificial apparatus to help observers improve their skills, French astronomer Charles Wolf leaped at the opportunity to investi-
75 76
77
lichkeit der wirklichen Erscheinung mit der Darstellung am Modell überrascht und fand das zu beobachtende Hauptmoment sehr sicher bestimmt.” Auwers, ed. Die Venus-Durchgänge. Vol. 1, 159. U.S. Transit of Venus Commission. Instructions. 41. Auwers, ed. Die Venus-Durchgänge. Vol. 1, 195; U.S. Transit of Venus Commission. Instructions. 49; Ministère de l'Instruction Publique et des Beaux-Arts. Conférence internationale. 30. Friedrich Wilhelm Bessel. Abhandlungen. Ed. Rudolf Engelmann. Leipzig: Verlag von Wilhelm Engelmann, 1875-1876. Vol. 3, 300-04.
Scientific Coordination as Ethos and Epistemology
321
gate the causes of personal equations and the means to eradicate them.78 He conducted a long series of experiments on himself and the staff of the Paris Observatory in 1863-64, using an apparatus of his own construction. At first his results seemed nothing less than spectacular. After three months of practicing with the machine, he was able to reduce his own absolute personal equation from 0.30 to 0.11 seconds. But after that, no further progress seemed possible, despite the most focused efforts of attention: In the present conditions, it is totally impossible for me to observe in a different way, and even the strictest attention does not allow me to seize the least dead time from the moment when I hear the second and that when I fix the position of the star.
Wolf drew radical conclusions from his own failure to eliminate error: Does this mean that time does not exist? Evidently not. By its very nature, in a good observer, it should be completely imperceptible; and it is even an indispensable condition of the consistency of the mode of observation of an astronomer: astronomy should not be perfectible.79
What had begun as a quest for perfection, of observation without error, attained through conscious practice, ended as the acceptance of unconscious and incorrigible lags in perception, so long as these were constant. Errors could not be entirely erased, but they could be made as predictable as the stars themselves. The role of the artificial transit machines in achieving constancy, if not perfection, was not trivial. Wolf’s results contradicted those of the Swiss astronomers Hirsch and Plantamour, who had found significant inter- and intra-observer variations in their absolute and relative personal equations.80 Central to resolving this issue, and indeed all issues concerning the personal equation, was a determination of the efficacy 78 79
80
Rudolphe Radau. “Sur les erreurs personnelles.” Moniteur Scientifique-Quesneville (November 15, 1865): 16-25. “Dans les conditions actuelles, il m’est tout à fait impossible d’observer autrement que je ne le fais, et l’attention la plus sévère ne me permet de saisir le moindre temps mort entre le moment où j’entends la seconde et celui où je pointe la position de l’étoile.” – “Est-ce à dire que ce temps n’existe pas? Évidemment non. Par sa nature même, chez un bon observateur, il doit être complètement imperceptible; et c’est même là une condition indispensable de la constance du mode d’observation d’un astronome: l’astronome ne doit pas être perfectible.” Charles Wolf. “Recherches sur l’équation personnelle dans les observations de passage.” Annales de l’Observatoire de Paris. Mémoires, 8 (1866): 188. Rudolphe Radau. “Sur les erreurs personnelles.” 22; Adolphe Hirsch. “Sur les corrections et équations personnelles dans les observations chronographiques de passage.” Bulletin de la Société des sciences naturelles de Neuchâtel 6 (1863): 365-72.
322
Peter Galison/Lorraine Daston
of training: how much could absolute personal equations be reduced by attention and education? To what extent was the irreducible error that remained a physiological fact about the observer’s constitution, to what extent a variable dependent on contingent circumstance? In a meticulous review of personal equation experiments to date, and against the background of the two recent transits of Venus, Lyon astronomer François Gonnessiat attempted in 1892 to sort out the contributing elements of the personal equation. The differences between the registration of visual, auditory, and tactile stimuli, the influence of the time of day, the finite speed of transmission of nervous impulses, the internalized time rhythms of seasoned observers – all were submitted to painstaking review and experiment. Gonnessiat’s aim was to classify the sources of error, and thereby to distinguish between “a necessary psychological error, and a different error depending more particularly on attention and capable of being reduced by training.” Errors would be sorted out into conscious and unconscious, voluntary and involuntary, psychological and physiological, subjective and objective. Indeed, the exercise of classification re-affirmed the boundary between the subjective and objective observer, between will on the one hand and nerves and muscles on the other. Kantian antinomies were remapped onto “erreurs rhythmiques” and “erreurs décimales.” Yet for all Gonnessiat’s care and thoroughness, the exercise failed: one component of the personal equation straddled the fence, precisely poised between subjective and objective, “l’erreur psycho-physiologique.” This was “the time that an optical impression needs to provoke the muscle play of the hand”– that is, the precise point of intersection between will and the world.81 In contrast to the map of the heavens or the exact shape of the earth, solar parallax was at least in principle a well-defined object. The devil lay in measuring it, since parallax by definition requires at least two observers positioned at either end of a baseline. Since the seventeenth century, astronomers had been convinced that the transits of Venus opened up the royal road for the determination of solar parallax; the scale and expense of the eighteenth and, especially, the nineteenth-century expeditions bear witness to the strength of their conviction. Yet even when the weather co-operated, even when colonial networks smoothed the way of the itinerant astronomers, even when observers practiced for 81
“. . . une erreur physiologique nécessaire, et une autre erreur dépendent plus particulièrement de l’attention et susceptible de réduction par l’exercice.” / “. . . le temps qui met une impression reçue par l’œil à déterminer le jeu des muscles de la main”. François Gonnessiat. Recherches sur l’équation personnelle dans les observations astronomiques de passage. Paris: G. Masson, 1892. 151.
Scientific Coordination as Ethos and Epistemology
323
months beforehand on identical artificial transit machines, the difficulties of coordination (even at the intra-individual level) pushed the precise determination of the parallax value out of reach. Conclusion: Peirce as Philosopher of Scientific Coordination For Peirce, the coordination of observers and the cancellation of scientific error became both the means toward and the metaphor of the bedrock reality sought by science. Peirce’s whole life was intertwined with the alignment of systems of observers. From immediately after his graduation from Harvard in 1859, when he signed on as a temporary aide at the U.S. Coast and Geodetic Society, to his separation from the Survey in 1891, Peirce embarked on a succession of field expeditions, theoretical inquiries into the coordination of instrumentation and statistical procedures, and international conferences designed to bring together measurements by different countries. Throughout his life, and across his several disciplinary bases, Peirce sought knowledge that would transcend what he called the “vagaries” of individualist inquiry. Before the Carte du Ciel was launched, Peirce was deeply engaged in Harvard’s own massive sky-mapping project.82 Between February 1872 and January 1875, he observed at a variety of East Coast sites, recording not only the brightness but also the color of stars.83 When Pickering protested in 1877 that Peirce’s magnitudes had fractional errors as high as 1/10, the philosopher-metrologist shot off an angry letter to his father, the Harvard mathematician Benjamin Peirce: This figure [of a fractional brightness error equal to 1/10] is not derived from the comparison of my observations among each other, but by the comparison of different observers among each other, so as to include all sources of error. 82
83
More precisely, Peirce employed a Zoellner astrophotometer, a device which allowed a star to be compared with an adjustable reference light, in order to determine stellar magnitudes. Victor F. Lenzen. “Charles Peirce as Astronomer.” Studies in the Philosophy of Charles Sanders Peirce. Ed. Edward C. More and Richard S. Robin. Amherst: University of Massachusetts Press, 1964. 44ff.; also Lenzen. “The Contributions of Charles S. Peirce to Metrology.” Proceedings of the American Philosophical Society 109 (1965): 29-46; Lenzen. “Charles S. Peirce as Mathematical Geodist.” Transactions of the Charles S. Peirce Society 8 (1972): 90-105; Lenzen. “The Role of Science in the Philosophy of C.S. Peirce.” Logik, Erkenntnis- und Wissenschaftstheorie, Sprachphilosophie, Ontologie und Metaphysik. Akten des XIV. Internationalen Kongresses für Philosophie, 2-9. September 1968. Vienna: Herder, 1969. 371-76.
324
Peter Galison/Lorraine Daston
2nd My probable error is smaller than that of any other observer . . . Our friend Pickering doesn’t know how to appreciate my work, that’s all.84
For Peirce, it went without saying that the error had to include an interobserver component that was combined with the intra-observer spread of results. His sense that the right result would only emerge from the community’s efforts is thus (unlike Pickering’s use of purely personal error) built into his expression of the probable error machinery itself. Peirce’s persistent concern, expressed in the details of his scientific work as well as the precepts of his philosophy, was to balance the idiosyncrasies and errors of single measurements and of individuals against one another in a great cancellation of errors. As he put it in 1868, The real, then, is that which, sooner or later, information and reasoning would finally result in and which is therefore independent of the vagaries of me and you. Thus the very origin of the conception of reality shows that this conception essentially involves the notion of a COMMUNITY, without definite limits, and capable of an indefinite increase of knowledge.85
Peirce’s daily activities with the Survey put him dead center within the new-style scientific community of his day, the coordinated activities of observer networks – his first Survey trip to Europe was from June to March 1871. After working at the Harvard Observatory some seven or eight months, Peirce composed a manuscript that built on his earlier concern with eschewing the “vagaries of me and you,” but went on to reflect the specific nature of observatory practice.86 Here he asked after the existence of “something independent of what you or I or any number of men, may think about it.” “What,” Peirce asked, “is the mode of existence of this reality?” By his lights, all right reasoning proceeds either from reason or from observation. Sometimes similar results issue from different premises, as in the various derivations of the rotation of the earth. But Peirce insisted that in fact all – not some – observations differ in this way as no two observers can ever have the “same” observation. “The observations which I made yesterday are not the same which I make today. Nor are simultaneous observations at different observatories the same, however close together the observatories are placed. Every man’s senses are his observatory.”87 Observation alone is incom84
85 86 87
Letter of Charles S. Peirce to Benjamin O. Peirce, 11 February 1877, MSS. Charles Sanders Peirce L33, Correspondence C.S. Peirce-B.O. Peirce, Houghton Library, Harvard University. Charles S. Peirce. “Consequences of Four Incapacities” [1868]. Writings. Vol. 2, 239. Charles S. Peirce, MS 204 (Fall 1872). “Chapter IV of Reality.” Writings. Vol. 4, 54. Ibid. Vol. 4, 55.
Scientific Coordination as Ethos and Epistemology
325
petent to contain the judgment that two observations are similar; that is a task for reason. By June 1873, Peirce was proposing to his father, then Superintendent of the Coast and Geodetic Survey, that he (Charles) undertake a world tour of measurement. The only places at which the length of the pendulum has actually been measured are London, Paris, Berlin & Switzerland, & the connection between these is nothing except Paris & London & there it might be better. In order therefore for me to connect the pendulum work on this continent with that all over the globe, I ought to take my apparatus to those places. It is a matter of the highest importance.88
Not only would he conduct measurements on the Sandwich Islands and other Pacific atolls, he would take the opportunity of his trip to join a transit of Venus expedition; from Japan and China he would head out across Egypt, swinging a pendulum from atop the great pyramid; from there he would sweep up to Europe, where he would compare instruments with his colleagues at the Europäische Gradmessung. Following the Paris International Geodetic Society meeting in September 1875, Peirce’s grand pendulum tours took him to Geneva, Paris, and Berlin, and culminated in his classic memoir on “Measurements of Gravity at Initial Stations in America and Europe” (1879). Even in this technical scientific text, two essential features of the ethos and epistemology of scientific coordination emerge clearly. First, Peirce argued in his cover letter to his father (then the Survey’s consulting geometer) that “the value of gravity-determinations depends upon their being bound together, each with all the others which have been made anywhere upon the earth.”89 Second, Peirce insisted that not only individuals but also nations must sacrifice their own interests and ambitions for the sake of the scientific community. This self-denial in the service of co-ordination, writ large, was the same moral principle that had led British astronomers to give up their superior reflecting telescopes in favor of the world stan88 89
Letter of Charles S. Peirce to Benjamin O. Peirce, 16 June 1873, MSS. C.S. Peirce L33, Houghton Library, Harvard University. Consequently, both Peirce senior and junior had supported making “relative” measurements of pendulum motion rather than attempting absolute measurements both of the pendulum arm and the time of oscillation. The idea behind their method was the following: A pendulum of length l, gravitational acceleration g, and period T obeys the relation T = ʌ (l/g)1/2. Therefore the ratio T1/T2 is just (g2/g1)1/2, since by using the same pendulum at each site, the length l drops out, leaving only the time T to be measured in order to fix g2/g1.
326
Peter Galison/Lorraine Daston
dard selected by the Carte du Ciel. National pride and ambition, Peirce argued, must yield to the larger network: At the meeting of the International Geodetic Congress in Paris, in 1875, it was resolved, at the suggestion of the writer, that the different states should carry their pendulums to Berlin and swing them in the Eichungsamt there. This has already been done by Switzerland and Austria, and will be done hereafter by every survey which is not willing to sacrifice the solution of a great problem to forms of action based on national exclusiveness. Geodesy is the one science the successful prosecution of which absolutely depends upon international solidarity.90
But what if both individuals and nations were incorrigible scientific egotists, à la Pickering? Peirce argued hard against the inevitability of individualism, and in “Grounds of Validity” (1869) explicitly opposed such selfishness to scientific objectivity: There is a psychological theory that man cannot act without a view to his own pleasure. This theory is based on a falsely assumed subjectivism. Upon our principles of the objectivity of knowledge, it could not be based; and if they are correct, it [the selfishness of men] is reduced to an absurdity.
For Peirce, the primacy of the community was a fact and not merely a pious hope, for he concluded from concern for those left behind after death and habitual identification with the communal destiny “that men do not make their personal interests their only ones, and therefore may, at least, subordinate them to the interests of the community.”91 The objective knower must be many-headed. It was thus reasonable to hope that humanity might voluntarily embrace the morality of “complete self-identification of one’s own interests with those of the community” that all inductive inference – and therefore most of science and life – in Peirce’s view presupposed.92 There is once again evidence that Peirce’s own experience in the collective conduct of observation and measurement had taught him that rationality imposed from above was futile. In an 1899 manuscript outlining the guiding principles of the U.S. Office of Weights and Measures, he observed that the full force of the Napoleonic empire had been insufficient to enforce the metric system in its dominions: In Prussia, for example, it was for some years a criminal offence to have a Rhenish foot rule in one’s possession; and yet I well remember that when I wished to have some carpentry done in one of the halls of the Berlin Eichungs90 91 92
Charles S. Peirce. “Measurements of Gravity.” 81. Charles S. Peirce. “Grounds of Validity” [1869]. Writings. Vol. 2, 271; emphasis added. Ibid. Vol. 2, 268-72.
Scientific Coordination as Ethos and Epistemology
327
amt, the mechanic whom I brought in for the purpose demanded that all the data I had to give him should be expressed in Rhine feet and inches.93
Peirce’s identification of a man’s senses with an observatory binds together two problem domains of utterly different scales. Like the establishment of results connecting individuals’ sense perceptions, linking astronomical observatories had come to stand for the search for realism, the overcoming of a nominalism that would separate us from our later selves, our selves from others, and one observatory from another. Forced coordination, however, would never work, and the futility of trying to fix belief by despotic means was a long-standing theme in Peirce’s more abstract writing. In the practical application, not only was the scientific method better suited to fixing belief than despotism; despotism was also thoroughly incompatible with American values, as Peirce went on to say: Americans would never consent to such restrictions upon their liberty. They would make a political issue of them if congress were ever so ill-advised as to enact them, which it never would do. The individual in this country expects to manage his own business in his own way. Our principle is to leave police matters, as much as possible, to the states; and nothing would be gained by an interference on the part of the general government with the police side of the work of metrology.94
Beyond expelling police power from science to the federal government, Peirce was arguing, in his technical work and in his philosophy, that a moral training for metrology was a necessary prerequisite to the science. For Peirce, the word “standard” resonated with a double meaning, at once defining weights and measures and the moral code which created and upheld those units.95 It was an internalized morality of conscience, rather than of obedience to external authority. In this same 1899 manuscript, Peirce wrote: “The office ought to be so administered as to create or maintain a distinct tradition of scientific and metrological morals within the office . . . Young men have to be trained from boyhood to the reverence for the laws of morals and honor which alone can give the best fruit.”96
93
94 95 96
Charles S. Peirce. “On the Proper Functions of a National Office of Weights and Measures” [1899]. Unpublished MS., MSS. Charles S. Peirce, Houghton Library, Harvard University. We thank Christian Kloesel and the Peirce Edition Project for the date of the manuscript. Charles S. Peirce, “MS 1899.” Writings. Vol. 5. For similar ambiguities concerning standards, cf. Schaffer. “Late Victorian Metrology and its Instrumentation.” Peirce. “Proper Functions.”
328
Peter Galison/Lorraine Daston
In his work on the Internationale Gradmessung Peirce had seen at first hand the need for more than a new branch of science; he had come to believe that a new kind of scientist was needed, one whose skills and morals both aimed towards the subjection of excessive willfulness. But if data, measuring missions, and instruments needed to be coordinated, so too did the great scientific and military organizations of the late nineteenth century. Between wars and sometimes straight through them, surveyors, astronomers, and physicists learned to make sacrifices, to hash out their differences until sufficient agreement could be obtained to cover the world with an interlaced grid of observers and data. Some looked down, to correct their maps, steer their boats, and correct their physical theories. Others looked up to measure the solar-system distances. All helped form a new kind of collective science with an internationalism made possible by the coordination of an ever more powerful nationalism. And with that coordination came the framework from which it became possible to bring planetary-scale objects into existence: the distribution of gravitational force, the detailed shape of the earth, the distribution of stars across the whole of the visible sky, and the absolute measure of the solar system.
WORKS CITED Académie des Sciences, Paris. Passage de Vénus du décembre 1882. Rapports préliminaires. Paris: Gauthier-Villars, 1883. Airy, George Biddell, ed. Account of Observations of the Transit of Venus, 1874, December 8, Made under the Authority of the British Government. And of the Reduction of the Observations. London: Her Majesty’s Stationery Office, 1881. Auwers, Arthur, ed. Die Venus-Durchgänge 1874 und 1882. Bericht über die Deutschen Beobachtungen. 6 vols. Berlin, 1887-98. Bacon, Francis. “New Atlantis” [1627]. Lord Bacon’s Works. Vol. 2. Ed. Basil Montagu. London: William Pickering, 1825-34. Baeyer, J.J. “Bericht über den allgemeinen Standpunkt der Preussischen Vermessungen in Bezug auf die mitteleuropäische Gradmessung, und im Besonderen über die im Jahre 1864 ausgeführten Arbeiten.” General-Bericht über die mitteleuropäische Gradmessung für das Jahr 1864. Berlin: Reimer, 1865. Bernard, Claude. Introduction to Experimental Medicine [1865]. Trans. Henry Copley Greene. New York: Dover, 1957. Bessel, Friedrich Wilhelm. Abhandlungen. 3 vols. Ed. Rudolf Engelmann. Leipzig: Verlag von Wilhelm Engelmann, 1875-1876. Biagioli, Mario. Galileo, Courtier. The Practice of Science in the Culture of Absolutism. Chicago and London: University of Chicago Press, 1995.
Scientific Coordination as Ethos and Epistemology
329
Bigg, Charlotte. “Photography and Labour History of Astrometry. The Carte du Ciel.” The Role of Visual Representations in Astronomy. History and Research Practice. Ed. Klaus Hentschel. Thun and Frankfurt a.M.: Deutsch, 2000. 90-106. Bonneuil, Christophe. “The Manufacture of Species. Kew Gardens, the Empire, and the Standardisation of Taxonomic Practices in Late Nineteenth-Century Botany.” Instruments, Travel and Science. Itineraries of Precision from the Seventeenth to the Twentieth Century. Ed. Marie-Noëlle Bourguet, Christian Licoppe, and H. Otto Sibum. London and New York: Routledge, 2002. 189-215. Bourguet, Marie-Noëlle. “Landscape with Numbers. Natural History, Travel and Instruments in the Late Eighteenth and Early Nineteenth Centuries.” Instruments, Travel and Science. Itineraries of Precision from the Seventeenth to the Twentieth Century. Ed. idem, Christian Licoppe, and H. Otto Sibum. London and New York: Routledge, 2002. 96-125. Brian, Éric. “Transactions statistiques au XIXe siècle. Mouvements internationaux de capitaux symboliques.” Actes de la recherche en sciences sociales 145 (décembre 2002): 34-46. Bronowski, Jacob. Science and Human Values [1956]. New York: Harper & Row, 1975. Cahan, David. An Institute for Empire. The Physikalisch-Technische Reichsanstalt, 18711918. Cambridge: Cambridge University Press, 1989. Canales, Jimena. “Photogenic Venus. The ‘Cinematographic Turn’ and Its Alternatives in Nineteenth-Century France.” Isis 93 (2002): 585-613. Canales, Jimena. Sensational Differences. Individuality in Observation, Experimentation, and Representation (France 1853-1895). Ph.D. Harvard University, 2003. Cannon, Susan Faye. “Humboldtian Science.” Science in Culture. The Early Victorian Period. New York: Science History Publications, 1978. Centralbureau der Europäischen Gradmessung. Verhandlungen der vom 13. bis 16. September 1880 zu München abgehaltenen sechsten allgemeinen Conferenz der Europäischen Gradmessung. Berlin: Reimer, 1881. Cooper, Alix. “The Household.” The Cambridge History of Early Modern Science. Ed. Katharine Park and Lorraine Daston. Cambridge: Cambridge University Press, 2006. 224-37. Cotreau, James D. The Historical Development of the Universal Postal Union and the Question of Membership. Boston: n. publ., 1975. Daston, Lorraine and Peter Galison. “The Image of Objectivity.” Representations 40 (1992): 81-128. Débarbat, Suzanne et al., eds. Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Dordrecht, Boston, and London: Kluwer, 1988. Descartes, René. “Discours de la méthode” [1637]. Œuvres de Descartes. Vol. 6. Ed. Charles Adam and Paul Tannery. Paris: Léopold Cerf, 1902. 1-78. Dettelbach, Michael. “Global Physics and Aesthetic Empire. Humboldt’s Physical Portrait of the Tropics.” Visions of Empire. Voyages, Botany, and Representations of Nature. Ed. Peter H. Reill and David Philip Miller. Cambridge: Cambridge University Press, 1996. 258-92. Eamon, William. Science and the Secrets of Nature. Books of Secrets in Medieval and Early Modern Culture. Princeton: Princeton University Press, 1994. Elzinga, Aant and Catharina Landström, eds. Internationalism in Science. London: Taylor Graham, 1996. Faye, Hervé. “Rapport sur le rôle de la photographie dans l’observation du passage de
330
Peter Galison/Lorraine Daston
Vénus.” Recueil des mémoires, rapports et documents relatifs à l’observation du passage de Vénus sur le soleil. Paris: Firmin Didot Frères, 1874. 227-36. Faye, Hervé. “Sur l’observation photographique des passages de Vénus et sur un appareil de M. Laussedat.” Recueil des mémoires, rapports et documents relatifs à l’observation du passage de Vénus sur le soleil. Paris: Firmin Didot Frères, 1874. 175-85. Findlen, Paula. Possessing Nature. Museums, Collecting, and Scientific Culture in Early Modern Italy. Berkeley and Los Angeles: University of California Press, 1994. Findlen, Paula. “Masculine Prerogatives. Gender, Space, and Knowledge in the Early Modern Museum.” The Architecture of Science. Ed. Peter Galison and Emily Thompson. Cambridge and London: MIT Press, 1999. 29-58. Flammarion, Camille. “La photographie céleste à l’Observatoire de Paris.” Revue d'Astronomie Populaire 5 (1886): 42-57. Flammarion, Camille. “Le Congrès astronomique pour la photographie du ciel.” Astronomie 6 (1887): 161-69. Fleming, James Rodger. Meteorology in America, 1800-1870. Baltimore and London: Johns Hopkins University Press, 1990. Forbes, George. The Transit of Venus. London and New York: Macmillan, 1874. Friedman, Robert Marc. Appropriating the Weather. Vilhelm Bjerknes and the Construction of a Modern Meteorology. Ithaca: Cornell University Press, 1989. Galison, Peter. Image and Logic. A Material Culture of Microphysics. Chicago and London: University of Chicago Press, 1997. Galison, Peter. Einstein’s Clocks, Poincaré’s Maps. Empires of Time. New York: W.W. Norton, 2003. General-Bericht über die mitteleuropäische Gradmessung für das Jahr 1862. Berlin: Reimer, 1863. Gonnessiat, François. Recherches sur l’équation personnelle dans les observations astronomiques de passage. Paris: G. Masson, 1892. Hahn, Roger. The Anatomy of a Scientific Institution. The Paris Academy of Sciences, 1666-1803. Berkeley and Los Angeles: University of California Press, 1971. Hannaway, Owen. “Laboratory Design and the Aim of Science. Andreas Libavius versus Tycho Brahe.” Isis 77 (1986): 585-610. Hellmann, Gustav. “Die Entwicklung der meteorologischen Beobachtungen in Deutschland von den ersten Anfängen bis zur Einrichtung staatlicher Beobachtungsnetze.” Abhandlungen der Preussischen Akademie der Wissenschaften, Physikalisch-Mathematische Klasse 1 (1926): 1-25. Hirsch, Adolphe. “Sur les corrections et équations personnelles dans les observations chronographiques de passage.” Bulletin de la Société des sciences naturelles de Neuchâtel 6 (1863): 365-72. Hoffleit, Dorrit. Some Firsts in Astronomical Photography. Cambridge: Harvard College Observatory, 1950. Hunter, Michael. The Royal Society and Its Fellows 1660-1700. Morphology of an Early Scientific Institution. Chalfont St. Giles: British Society for the History of Science, 1982. Institut de France-Académie des Sciences, ed. Congrès astrophotographique international tenu à l’Observatoire de Paris pour le levé de la Carte du Ciel. Paris: GauthierVillars, 1887. Kristensen, Leif Kahl. “T.N. Thiele and the Carte du Ciel.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 59-63.
Scientific Coordination as Ethos and Epistemology
331
Lankford, John. “Amateurs and Astrophysics. A Neglected Aspect in the Development of a Scientific Specialty.” Social Studies of Science 11 (1981): 275-303. Lankford, John. “The Impact of Photography on Astronomy.” Astrophysics and Twentieth-Century Astronomy to 1950. Ed. Owen Gingerich. Cambridge and New York: Cambridge University Press, 1984. 16-39. Lankford, John. “Photography and the Nineteenth-Century Transits of Venus.” Technology and Culture 28 (1987): 648-57. Leibniz, Gottfried Wilhelm. Untitled fragment [1677]. Die philosophischen Schriften von Gottfried Wilhelm Leibniz. Ed. Carl Immanuel Gerhardt. Berlin: Weidmann, 1875-90. Vol. 7, 184-89. Lenoir, Timothy. Instituting Science. The Cultural Production of Scientific Disciplines. Stanford: Stanford University Press, 1997. Lenzen, Victor F. “Charles Peirce as Astronomer.” Studies in the Philosophy of Charles Sanders Peirce. Ed. Edward C. More and Richard S. Robin. Amherst: University of Massachusetts Press, 1964. 33-50. Lenzen, Victor F. “The Contributions of Charles S. Peirce to Metrology.” Proceedings of the American Philosophical Society 109 (1965): 29-46. Lenzen, Victor F. “The Role of Science in the Philosophy of C.S. Peirce.” Logik, Erkenntnis- und Wissenschaftstheorie, Sprachphilosophie, Ontologie und Metaphysik. Akten des XIV. Internationalen Kongresses für Philosophie, 2-9. September 1968. Vienna: Herder, 1969. 371-76. Lenzen, Victor F. “Charles S. Peirce as Mathematical Geodist.” Transactions of the Charles S. Peirce Society 8 (1972): 90-105. Mill, John Stuart. “The Spirit of the Age” [1831]. Collected Works of John Stuart Mill. Ed. John M. Robson et al. Toronto: University of Toronto Press, 1981-91. Vol. 32, 227-316. Ministère de l’Instruction Publique et des Beaux-Arts. Conférence internationale du passage de Vénus. Procès-Verbaux. Paris: Imprimerie Nationale, 1881. Moran, Bruce T. The Alchemical World of the German Court. Occult Philosophy and Chemical Medicine in the Circle of Moritz of Hessen, 1572-1632. Stuttgart: F. Steiner Verlag, 1991. Mouchez, Ernest B. La Photographie astronomique à l’Observatoire de Paris et la Carte du Ciel. Paris: Gauthier-Villars, 1887. Norman, Daniel. “The Development of Astrophotography.” Osiris 5 (1938): 560-94. O’Hora, Nathy P. “Astrographic Catalogues of British Observatories.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 135-38. Peirce, Charles S. “Grounds of Validity” [1869]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 2, 242-72. Peirce, Charles S. MS 204 [Fall 1872]. “Chapter IV of Reality.” Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 54-59. Peirce, Charles S. “Measurements of Gravity at Initial Stations in America and Europe.” [1879]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 79-144. Peirce, Charles S. “Six Reasons for the Prosecution of Pendulum Experiments” [1882]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 359-60.
332
Peter Galison/Lorraine Daston
Peirce, Charles S. “General Remarks upon Gravity Determinations, by John Herschel.” Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 365-68. Peirce, Charles S. “On the Proper Functions of a National Office of Weights and Measures” [1899]. Unpublished MS., MSS. Charles S. Peirce, Houghton Library, Harvard University. Peirce, Charles S. “Consequences of Four Incapacities” [1868]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 2, 239. Radau, Rudolphe. “Sur les erreurs personnelles.” Moniteur Scientifique-Quesneville (15 November 1865): 22. Schaffer, Simon. “Late Victorian Metrology and Its Instrumentation. A Manufactory of Ohms.” Invisible Connections. Instruments, Institutions, and Science. Ed. Robert Bud and Susan E. Cozzens. Washington: Spie Optical Engineering Press, 1992. 2356. Schaffer, Simon. “Metrology, Metrification and Victorian Values.” Victorian Science in Context. Ed. Bernard Lightman. Chicago and London: University of Chicago Press, 1997. 438-74. Scheiner, Julius. Die Photographie der Gestirne. Leipzig: Wilhelm Engelmann, 1897. Schiebinger, Londa. The Mind Has No Sex? Women in the Origins of Modern Science. Cambridge and London: Harvard University Press, 1989. Sellers, David. The Transit of Venus. The Quest to Find the True Distance to the Sun. Leeds: Maga Velda Press, 2001. Shapin, Steven. “The House of Experiment in Seventeenth-Century England.” Isis 79 (1988): 373-404. Stagl, Justin. A History of Curiosity. The Theory of Travel, 1550-1800. Chur: Harwood, 1995. Standage, Tom. The Victorian Internet. The Remarkable Story of the Telegraph and the Nineteenth-Century’s On-Line Pioneers. New York: Walker, 1998. Turner, Herbert Hall. The Great Star Map. New York: E.P. Dutton, 1912. U.S. Transit of Venus Commission. Instructions for Observing the Transit of Venus, December 6, 1882. Washington, D.C.: Government Printing Office, 1882. Verhandlungen der vom 20. bis 29. September 1875 in Paris vereinigten Permanenten Commission der Europäische Gradmessun. Berlin: Reimer, 1875. Völter, Ulrich. Geschichte und Bedeutung der internationalen Erdmessung. Munich: Verlag der Bayerischen Akademie der Wissenschaften, 1963. Wayman, Patrick A. “The Grubb Astrographic Telescopes.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 139-42. Weimer, Théo. “Naissance et développement de la Carte du Ciel.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 29-32. White, Graeme L. “The Carte du Ciel – The Australian Connection.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 45-51. Winterhalter, Albert G. The International Astrophotographical Congress and A Visit to
Scientific Coordination as Ethos and Epistemology
333
Certain European Observatories and other Institutions. Report to the Superintendent. Washington: Government Printing Office, 1889. Wise, Norton and Crosbie Smith. Energy for Empire. A Biographical Study of Lord Kelvin. Cambridge: Cambridge University Press, 1989. Wolf, Charles. “Recherches sur l’équation personnelle dans les observations de passage.” Annales de l’Observatoire de Paris. Mémoires 8 (1866): 188. Woolf, Harry. The Transits of Venus. A Study in Eighteenth-Century Science. Princeton: Princeton University Press, 1959.
STEFAN DITZEN
Breaking, Grinding, Burning: Instrumental Aspects in Early Microscopical Pictures 1. On the Importance of Lenses and the Respect for Craftsmanship An early example of the use of a lens can be found in the cross of Henry1 from the Basel Cathedral treasury, kept today in the Museum for Arts and Crafts Berlin.2 It is composed of a wooden core covered with golden and silver sheets. The central cameo is surrounded by four big, semi-convex quartz crystals, each of them on one arm of the cross. Beneath the left and the right stone the relics of the Emperor and Saint Henry II (1002-1024) are located. This is indicated by small slips of parchment that can be read through the crystallized stones. The other stones cover particles of wood that are supposed to have belonged to the cross of Jesus. These pieces are arranged in the form of a cross.3 Additionally, the upper stone covers another relic, possibly of Henry’s wife, Saint Kunigunde. The stones have more than a decorative function. Due to their clarity and semi-convex shape, they serve to magnify the relics beneath them. The small relics are more visible and the writings on the parchment are easier to read. This kind of application of crystals can be found with other relics too, enlarging the hair of Mary or other fragments of the cross. Therefore, one can assume that the properties of such lenses were used even at the time of the Carolingians in order to enlarge very small relics.4 In this religious context, the significance of the shaped stone 1 2 3
4
“Heinrichskreuz.” Cf. the catalogues of the Museum for Arts and Crafts Berlin (1963). Vol. I, no. 3 (inventory number 17,79). The cross under the upper stone is held by two angels. This kind of presentation goes back to the time around 1400. Consequently, it is difficult to date single aspects of this presentation. The appearance of the cross cannot be said to be the original shape since it was altered several times in the fourteenth and fifteenth centuries. Heinz Herbert Mann. Augenglas und Perspektiv. Studien zur Ikonographie zweier
Breaking, Grinding, Burning
335
changed from a decorative element to a visual aid. The metaphorical use of optical assistance in connection with religious ideas was obvious at least by the time of Nicolaus von Cues’s (1401-1464) De beryllo. Cusanus understood his Christian doctrine of coincidence as a pair of glasses that helped reason to “see” what remained closed to the human intellect.5 With the shift from the use of quartz crystals to the production of glass lenses some of these religious connotations were transferred as well. This is strongly suggested by the presentation of the glass lenses belonging to one of Galileo Galilei’s (1564-1642) telescopes, kept today in the collection of the Science Museum in Florence (Museo di Storia della Scienza). The broken lens6 of 30 mm in diameter is presented as a precious object in a baroque frame made of wood and ivory (fig. 1). This frame of 41 x 30 cm was made by Vittorio Croster in 1677.7 Its ivory carvings show a network of plants holding a variety of scientific instruments and a model of the cosmos. An inscription carved in ivory indicates that the frame holds the lens with which Galileo Galilei had been able to observe the moons of Jupiter.8 Due to the current manner of presentation, the visitor to the museum cannot observe anything other than a solid golden background underneath the lens. The optical assistance is not meant to focus on any observable object. On the contrary, the framing focuses the spectator’s view onto the lens itself. The tiny piece of glass is the exclusive object of interest. Hence, the presentation displays a precious construction un-
5
6
7
8
Bildmotive. Berlin: Gebr. Mann, 1992. 24f. Compare, for example, the use of such stones for a relic in the form of a disc in the cathedral treasury in Aachen. Cf. Kurt Flasch. Nikolaus von Kues. Die Idee der Koinzidenz. Quoted in Josef Speck, ed. Grundprobleme der großen Philosophen. Philosophie des Altertums und des Mittelalters. 5th ed. Göttingen: Vandenhoeck & Ruprecht, 2001. The inventory (no. 826) explains: “Un vetro obiettivo che fu già del celebre Galileo Galilei destinato in dono dal Serenissimo Granduca di Toscana Ferdinando II, per mezzo del quale scoperse tutte le novità celesti, e fra l’altre i quattro Pianeti intorno al corpo di Giove chiamati da esso Pianeti Medicei; il quale obiettivo vivente il medesimo Galilei accidentalmente si ruppe.” Mara Miniati, ed. Museo di Storia della Scienza. Catalogo. [Exhibit. cat.] Florence: Olschki, 1991. 60. “Nel luglio del 1677 da Vittorio Croster, intagliatore in avorio, fu fatta l’artistica cornice che tuttora racchiude il prezioso cimelio, il quale passò al Museo di Fisica il 18 maggio 1793 in ordine al Sovrano Rescritto del 19 aprile dello stesso anno.” Leo S. Olschki, ed. Catalogo degli strumenti del museo di storia della scienza. Florence: Olschki, 1954. “Coelum Linceae Galilei menti apertum Vitrea prima hac mole nondum visa ostendit sydera medicea iure ab inventore dicta. Sapiens nempe dominatrur et astris.” Miniati. Museo di Storia. 60.
336
Stefan Ditzen
Fig. 1: Broken lens of one of Galileo Galilei’s telescopes presented in a frame by Vittorio Croster.
derlining the value of this special piece of glass within the context of a scientific discovery. This arrangement reflects not only respect for Galilei’s work, but also the fact that his discovery could not have been made by the use of just any glass lens. Taken together, these considerations reinforce the view that respect for the precious stone has become respect for the craftsmanship that produced such glass lenses. The piece of groundglass itself became the relic to be admired. However, what made such a piece of glass so valuable in the seventeenth and eighteenth centuries? In connection with this question one needs to investigate the circumstances of the production and the use of lenses at that time.
Breaking, Grinding, Burning
337
2. The Production of Lenses and Problems of Instrumental Sight In the preface to Christian Gottlieb Hertel’s (1683-1743) Anweisung zum Glass-Schleiffen, the demand for excellent lenses for scientific research is reported: The ground glasses have contributed a lot to the precise knowledge of nature. To them we owe the fact that we know today a lot of things concealed from our predecessors, and the correctness of which one should not doubt, about which our predecessors only had speculations . . . But it should be well noted that the ground glasses need to have their right perfection if we want to obtain the use that we so praise. Otherwise they could obstruct the things or not really disclose them and therefore lead more to mistakes and illusions than to honest truth . . . For this reason professor Hertel did his honorable work of describing so clearly the Art of grinding glasses and of composing a diversity of useful instruments with ground glasses, and revealed with sincerity the skills he learned from his experience for everybody’s benefit, something that other people who are envious and addicted to profit would have kept secret.9
Due to the dimensions, it was difficult to reach the required “right perfection” for the lenses of microscopes. Furthermore, the problem of spherical and chromatic aberration10 resulting from the combination of at least two lenses within one single instrument led to even more confusion. This fact entailed the danger of optical deceptions and therefore artifacts within pictorial representations. On the other hand, the required degree of enlargement was easier to achieve by the use of multiple lenses. They could be of lesser curvature because the power of the lenses is multiplied through the combinations. This means the small single lens of a simple microscope had to be of a much greater curvature in order to achieve the necessary focal distance.11 In consequence, such instruments 9 10
11
Preface written by Hoff-Rath Wolff. Christian Gottlieb Hertel. Vollständige Anweisung zum Glass-Schleiffen . . . Halle, 1716. Chromatic aberration results from the shape of the lens being basically similar to a combination of prisms. The passing light is separated into the colors of the spectrum. Therefore microscopic images have colored edges and modifications. Spherical aberration blurs the images. The effect derives from the rays of light being more refracted at the edges of the lens than in the center. Consequently, the focal points of the light rays differ slightly. The best-preserved seventeenth century lens was produced by Antoni van Leeuwenhoek (1632-1723). It is only 1.2 mm in size and the radius of its curvature is approximately 0.7 mm on each side. The magnification is x270, and opinions differ as to whether this instrument can be regarded as representative of Leeuwenhoek’s microscopes or not. The possible magnification of such instruments has been speculated to be as much as x770. Cf. Edward G. Ruestow. The Microscope in the Dutch Republic. The Shaping of Discovery. Cambridge: Cambridge University Press, 1996.
338
Stefan Ditzen
were rather small and extremely difficult to handle. Additionally, the small round lenses absorbed a lot of light and made observation problematic, as Denicke remarked: Experience has shown that the small spheres are very unpleasant and difficult to use and to show especially the smallest parts of an object, because they let very little light through; yes, their power of magnification leads more to errors than to the detection of the truth, due to the lack of clarity and because of their tiring effect on the eyes.12
In all, observation with both kinds of instruments was a challenge for scientists. The compound as well as the simple microscope could possibly cause errors of representation. However, how were such fine biconvex lenses produced in the seventeenth century? The circumstances are documented within the textbooks on the “Art of Grinding Lenses,” written by Johannes Zahn (1641-1707), Johann Georg Leutmann (1667-1736) or the aforementioned Hertel.13 According to these works, the lenses could have been ground, melted, or blown in order to approach the “right perfection.”14 Hertel described in detail the process of producing the very small pieces of glass used as raw material for the lenses. Melted glass was pulled until it became long threads that were easy to break into pieces. These little fragments were picked up by the aid of a moistened pin and melted in a candle flame. The fragments assumed the shape of small balls as a result of this procedure. Another procedure was to melt a piece of glass into a mould of coal making use of a soldering instrument, which accelerated the process by adding more oxygen. A sphere of glass which was part of the soldering instrument removed the moisture from the breath of the blower (fig. 2). For the blowing of lenses glass tubes were closed by 12 13
14
C.L. Denicke. Vollständiges Lehrgebäude der ganzen Optik oder der Sehe-, Spiegel- und Strahlbrech-Kunst . . . Altona, 1757. 329f. Hertel. Vollständige Anweisung; Johannes Zahn. Oculus artificialis teledioptricus. Herbipoli, 1685; idem. Oculus artificialis teledioptricus. Nuremberg, 1702; Johann Georg Leutmann. Neue Anmerckungen vom Glaßschleiffen, darinnen die rechten Maschinen die Gläßer durch Hülffe dreyer Bewegungen zu mehrerer Vollkommenheit zubringen . . . Wittenberg, 1719; and Denicke. Vollständiges Lehrgebäude der ganzen Optik. J. van Zuylen discovered that most of Leeuwenhoek’s lenses were ground and polished. Only the very best ones examined had been blown. J. van Zuylen. “The Microscopes of Antoni van Leeuwenhoek.” Journal of Microscopy 121.3 (1981): 309-28. The lens produced by blowing glass is similar to the drop of water on the lower side of a soap bubble. Even the drop of glass of a light bulb used for flashlights can be used for magnifying and gives quite good results. Klaus Meyer. Geheimnisse des Antoni von Leeuwenhoek. Ein Beitrag zur Frühgeschichte der Mikroskopie. Lengerich: Pabst, 1998. 620ff.
Breaking, Grinding, Burning
339
Fig. 2: Procedure of melting a lens in the form of a little ball (fig. 5) as well as melting with the aid of a soldering tube (fig. 4).
melting one end. When such a tube was blown, the sphere on the end had a small bulge. This bulge, which could be used as a lens, was separated off by breaking the sphere.15 The most famous lenses of the seventeenth and eighteenth centuries were produced by Antoni van Leeuwenhoek (1632-1723). Most of these lenses were ground but it remained a secret how it had been possible for Leeuwenhoek to manufacture these rare optical wonders. The machines for grinding glass were big wooden constructions with either a rotating wheel or specially prepared bowls to grind the blank which was attached to a holding stick by a little bit of pitch.16 The bowls were made of lead, tin, iron, copper, or brass. As abrasives for the polishing the opticians used sands, ashes of tin, polishing red, and Tripel.17 It was absolutely necessary to clean the machine frequently after each working cycle, because a single grain of sand from the previous step could 15 16
17
van Zuylen. “Microscopes.” Denicke. Vollständiges Lehrgebäude der ganzen Optik. 270f. For the production of the smallest lenses and the procedure of blowing glass to produce lenses, cf. also Leutmann. Neue Anmerckungen. 38ff. Powder made out of the shells of diatoms.
340
Stefan Ditzen
scratch the lens and destroy the whole work.18 Both, the lens on the stick and the wheel or bowl were rotated during the procedure, while the stick with the lens had to be put in the position of different angles inside the bowl to obtain a regular and symmetrical grinding. With a perfectly vertical positioning the parts on the edge of the lens would have been ground more intensively than the part in the middle. By looking at the proportions of the grinding machines in different textbooks one can understand what made the finest lenses of sometimes only one millimeter in size so special and precious.19 The skills of the opticians played a decisive role in early microscopy and defined what became visible by the new instruments (This is a very similar effect to the production of special tips and slides for Scanning Tunneling Microscopy. Compare the article by Jochen Hennig in this book). Overall, one finds that in the esteem for the scientific discoveries, not only the lens of glass was ennobled, like in the case of Galilei’s lens, but also the whole process of production and therefore the machines for grinding were highly appreciated. At least this is suggested by the precious splendid machine of the mid-eighteenth century, which are kept in Florence.20
18
19
20
Cf. Hertel. Vollständige Anweisung. 6f. and 14f.; Denicke. Vollständiges Lehrgebäude der ganzen Optik. Part III, chapter 5ff.; furthermore Meyer. Geheimnisse des Antoni von Leeuwenhoek. 623. In the nineteenth century the value of the material was added to the worth of the lens due to the circumstances of production. Some opticians again produced lenses out of precious stones, but this has more to do with an attempt to improve the optical properties than with the symbolic meaning of the materials. Andrew Pritchard (1804-1882) tried in 1824 to produce lenses out of diamonds, but David Brewster (1781-1868) had already had the idea to do so in 1813. However, the very hard material was so difficult to grind and polish that such lenses never went into production. Moreover, the light was refracted by these lenses twice and therefore the image was doubled. It was much easier to produce lenses out of sapphires and therefore this stone was used more often. But by the 1830s production was more or less stopped due to Joseph Jackson Lister’s (1786-1869) discovery of the possibility of producing achromatic lens systems. They replaced the very expensive lenses made of precious stone. Gerard L’E. Turner. Essays on the History of the Microscope. Oxford: Senecio Publications, 1980. Probably the machine held a status between a collection piece that expressed the principles of lens production on the one hand and a practical use at court on the other hand, similar to the machines used for wood turning.
Breaking, Grinding, Burning
341
3. Microscopical Atlases and the Role of Pictorial Tradition In fact, not all microscopists at that time owned high quality lenses like the ones Leeuwenhoek used. He did not only keep the secret about the circumstances of their production but also did not sell any of his microscopes. Thus, with the questionable quality of lenses it remained doubtful whether all of the observed structures and colors belonged to the specimen or if they were caused by optical effects due to the apparatus. Moreover, a drawing of the observed structures produced by memory could be influenced by the imagination. This is the reason why drawing a microscopical view was supposed to take place under the circumstances of the shortest possible memorization. Many microscopists observed the specimen with one eye at the ocular, while they checked the drawing with the other eye at the same time. This procedure was complicated and required a lot of practice. Furthermore, it was only possible to focus on just a small part of the microscopical objects. This means drawing a microscopical picture was similar to scanning the observed structure step by step and putting it together in one single picture like a puzzle. What were the strategies used to avoid errors of misinterpretation resulting from both the lesser quality of lenses and the difficulties in practice? Information and pictures presented by other microscopists were used by scientists to confirm their own observations. In this relation the pictures seen formerly had an influence on further observations, on imagination and depictions. Especially for the plates of Robert Hooke’s (1635-1703) Micrographia21 one can find several transfers. These pictures were transmitted over time into several books either as a copy, a pictorial quotation, or a kind of imaginative transfer. This was due to the high quality of his microscopical viewing. The lesser the reliability of one’s own instrument was, and the weaker one’s own skills in depicting microscopical structures were, the more plausible and useful it seemed just to reproduce Hooke’s pictorial predecessors.22 With all their 21
22
Robert Hooke. Micrographia or Some Physiological Descriptions of Minute Bodies Made by Magnifying Glasses with Observations and Inquiries there upon. London, 1665. Even detailed observations on various microscopical objects, mainly insects, had already existed for some time before Hooke’s publication, but most of them were communicated by texts and not by pictures. Nevertheless, such texts, with their fantastic descriptions, defined something like a canon, a complete research program. Many of Hooke’s specimens can be found as early as Thomas Moffett’s (15531604) Theatrum insectorum (1634) or Henry Power’s (1623-1668) Experimental Philosophy (1664), the textual forerunner of Hooke’s Micrographia.
342
Stefan Ditzen
Fig. 3: Comparison of the representation of a louse by Robert Hooke (left) and by Johann Frantz Griendel (right).
details, their perfection and beauty, Hooke’s pictures overshadowed the capabilities of many other microscopists. On their way from book to book they reached an iconic status of the highest reliability, a claim for being more truthful than other depictions. Johann Frantz Griendel von Ach’s (1631-1687) Micrographia nova of 1687 shows what the results of scientists might have been if they had tried to avoid such a transfer. In contradiction to the title, the book did not offer a new type of micrography but primarily recapitulated Hooke’s observations. Contrary to other microscopists, Griendel did not copy the plates of the Micrographia but tried to depict his very own observations, although he was less experienced in doing so. The drawing gives the impression of uncertainty about the real appearance. There are no clear details to be found. For example, a structural differentiation between the leg and the feeler of a louse is completely missing (see the comparison in fig. 3: on the left side is the louse depicted by Hooke and on the right side the one by Griendel). It remains unclear whether the lesser quality of the result is due to the weak resolution of the instrument or to a lack of artistic talent.
Breaking, Grinding, Burning
343
In contrast, other scientists did use Hooke’s plates for their own publications. A comparison of Philippo Bonanni’s (1638-1725) Micrographia curiosa23 with Hooke’s atlas reveals the picture of the louse to be a copy.24 Hooke’s picture shows the louse clinging to a hair. This representation was probably chosen to give a scale with the animal. Another reason might have been to make reference to the creature’s size and living space. In contrast to this element, a shadow was added to the picture that shows the animal to be lying on a white background, maybe a sheet of paper. The louse of Bonanni likewise clings with its claws to a hair, but this time the inclination of the hair is just inverted. The body of the louse itself is quite similar to Hooke’s original and only the positions of the legs differ due to the new direction of the hair. While the shadow behind the animal is missing in Bonanni’s representation, even the different length of the feelers had been transferred (compare fig. 4: the louse in Bonanni’s book being on the upper right side and Hooke’s original version on the upper left). A comparison of this picture with other representations from the same publication makes obvious why Bonanni might have decided to use a copy. His very own skills in depicting microscopical observations were rather poor. The engravings of his observations are very schematic and show strongly reduced shapes similar to those of Griendel. This strongly suggests that the engraver had not been the factor of limitation. He had been able to engrave much more precisely than Bonanni’s drawings required. Consequently, Bonanni gave him the much more detailed picture of Hooke’s for the engraving of the louse, rather than his own drawings.25 Such a procedure was repeated several times in the following century. With his Observations, published in 1754, the French microscopist 23 24
25
Philippo Bonanni. Observationes circa viventia, que in rebus non viventibus reperiuntur. Cum Micrographia curiosa. Rome, 1699. Bonanni copied several motives from Hooke’s Micrographia. In some of the pictures, it becomes obvious how detailed the originals had been transferred, regardless of whether the view was supposed to represent a general type or a unique specimen. This can be seen, for example, in the copy of a fly’s eye that Bonanni transferred, in that he also copied the reflections of windows on the shining and reflecting surface of the compound eye. The precision of Hooke’s pictures even made the identification of one of the original objects possible. E.T. Tozer compared Hooke’s picture of a fossil shell, published in the Posthumous Works in 1705, with the collection of the British Museum and identified the original object. Cf. E.T. Tozer. “Discovery of an Ammonoid Specimen described by Robert Hooke.” Notes and Records of the Royal Society of London 44.1 (1990): 3-12.
344
Stefan Ditzen
Fig. 4: Comparison of the representation of a louse by Robert Hooke (upper left), Philippo Bonanni (upper right), Louis Joblot (lower left), and Johann Jacob Scheuchzer (lower right).
Louis Joblot (1645-1723) presented on the very first of his plates just the louse from Hooke’s Micrographia clinging to a hair. The picture is accompanied by other copies like, for instance, the representation of a bee’s sting on the very same plate (fig. 4 on the lower left). Joblot refers to the figure as being engraved exactly like Hooke’s version.26 26
“En voici la figure, que j’ai fait graver exactement de la même manière que M. Hook l’a dessinée afin de montrer précisément comme on la voit avec le Micro-
Breaking, Grinding, Burning
345
The same louse on the hair had even been transferred before, namely between 1731 and 1735 to Johann Jacob Scheuchzer’s (1672-1733) Physica sacra.27 Aside from Scheuchzer’s changes in representation, the surrounding frame, and the addition of further pictures, the transfer of Hooke’s original is still obvious (fig. 4 on the lower right). The animal was now holding a stretched hair as if it were a lance. Within the text Scheuchzer named Hooke and Leeuwenhoek among others as the sources of his work and thereby made clear his broad knowledge about the state of the art in microscopical sciences.28 Only the context had changed. The physico-theologian argued for a new interpretation of the Bible with the pictures. Together with Jan Swammerdam’s (1637-1680) view of a mosquito,29 the louse was interpreted as one of the biblical plagues. God had chosen little insects in contrast to big mammals to punish mankind because this meant an additional humiliation.30 So the context of such pictorial representations changed while the iconic matrix remained rather constant. Along with such chains of representations, the pictorial form of these “icons of microcosm” proved to be extremely constant for a number of microscopical representations, such as for fleas and mites as well. The Scheme XXXVI of the Micrographia shows two views of a mite, seen from a frontal position and at an angle. One can find a rather precise copy of this second version in the following centuries.31 The significance of the expression “icons of microcosm” is made clear by the front side of an advertisement produced in the nineteenth century. The paper promoted a microscopic projection by means of a
27
28 29 30
31
scope; & ce qui me l’a fait particulièrement choisir entre plusieurs autres, c’est qu’étant la plus grande de toutes celles qui sont dans ce livre, & ayant plus d’un demi-pié de longueur, elle fait mieux voir que les autres jusqu’à quel point cet instrument peut grossir les objets.” Louis Joblot. Observations d’histoire naturelle, faites avec le microscope, sur un grand nombre d’Insectes . . . Paris, 1754. 4. Johann Jacob Scheuchzer. Kupfer-Bibel, In welcher Die Physica Sacra, Oder Geheiligte Natur-Wissenschafft Derer In Heil. Schrifft vorkommenden Natürlichen Sachen, Deutlich erklärt und bewährt . . . Augsburg and Ulm, 1731-1735. Epist. P. 701; in Scheuchzer. Physica sacra. Vol. I, 175. Scheuchzer. Physica sacra. Tab. CXXVI. The author explicitly mentions Jan Swammerdam (1637-1680) as one of his sources (Hist. Insect. Pag. 95; in ibid. 173). “Es hat aber Gott dem Herrn in Bestraff- und Demüthigung der Egypter besonders gefallen sich solcher geringen und nichtsgeachteten Thierlein zu bedienen, als die Schnacken und Läuse sind, keines Wegs aber der Bären, Löwen, Pardeln, Schlangen, Crocodilen, davon letztere in dem Nil-Strom ohne diß vorhanden waren.” Scheuchzer. Physica sacra. 176. Cf. for instance Joblot’s publication of 1754 where he also showed versions of the animal pictured by Bonanni and Griendel.
346
Stefan Ditzen
Fig. 5: Picture of a mite by Robert Hooke (left) and on a poster for a microscopic exhibition (right).
solar microscope. The picture that advertised this Microscopic Exhibition was a highly standardized and widely known picture of a mite, namely Hooke’s view from the Micrographia (fig. 5: compare the mite from Hooke’s book on the left with the one from the poster32). Just as much as with the other pictures, there is obviously more than coincidental uniformity. This picture completely changed the context in comparison with Bonanni’s and Joblot’s use for their scientific textbooks. But it becomes evident how constant such a model of the seventeenth century could be. The advertisement comes from the time around 1835. This means that this early representation of a mite was still the picture of a mite as such about 170 years later.
32
Gerard L’E. Turner. Mikroskope. Munich: Callwey Verlag, 1981. 99; Emil-Heinz Schmitz. Das Mikroskop (= Handbuch zur Geschichte der Optik, Ergänzungsbd. II). Bonn: Wayenborgh, 1989. 354. Today the poster is in the archives of the Whipple Museum of the History of Science in Cambridge.
Breaking, Grinding, Burning
347
WORKS CITED Bonanni, Philippo. Observationes circa viventia, que in rebus non viventibus reperiuntur. Cum Micrographia curiosa. Rome, 1699. Denicke, C.L. Vollständiges Lehrgebäude der ganzen Optik oder der Sehe-, Spiegelund Strahlbrech-Kunst . . . Altona, 1757. Diedrichs, Christof. Vom Glauben zum Sehen. Die Sichtbarkeit der Reliquie im Reliquiar. Ein Beitrag zur Geschichte des Sehens. Berlin: Weißensee-Verlag, 2001. Griendel, Johann Frantz. Micrographia nova. Nuremberg, 1687. Hertel, Christian Gottlieb. Vollständige Anweisung zum Glass-Schleiffen . . . Halle, 1716. Hooke, Robert. Micrographia or Some Physiological Descriptions of Minute Bodies Made by Magnifying Glasses with Observations and Inquiries there upon. London, 1665. Joblot, Louis. Observations d’histoire naturelle, faites avec le microscope, sur un grand nombre d’Insectes . . . Paris, 1754. Leutmann, Johann Georg. Neue Anmerckungen vom Glaßschleiffen, darinnen die rechten Maschinen die Gläßer durch Hülffe dreyer Bewegungen zu mehrerer Vollkommenheit zubringen . . . Wittenberg, 1719. Mann, Heinz Herbert. Augenglas und Perspektiv. Studien zur Ikonographie zweier Bildmotive. Berlin: Gebr. Mann, 1992. Meyer, Klaus. Geheimnisse des Antoni von Leeuwenhoek. Ein Beitrag zur Frühgeschichte der Mikroskopie. Lengerich: Pabst, 1998. Miniati, Mara, ed. Museo di Storia della Scienza. Catalogo. [Exhibit. cat.] Florence: Olschki, 1991. Moffett, Thomas. Insectorvm sive minimorum animalivm theatrum. London, 1634. Olschki, Leo S., ed. Catalogo degli strumenti del museo di storia della scienza. Florence: Olschki, 1954. Petri, Richard Julius. Das Mikroskop. Von seinen Anfängen bis zur jetzigen Vervollkommnung für alle Freunde dieses Instruments. Berlin: Schoetz, 1896. Power, Henry. Experimental Philosophy. London, 1664. Ruestow, Edward G. The Microscope in the Dutch Republic. The Shaping of Discovery. Cambridge: Cambridge University Press, 1996. Scheuchzer, Johann Jacob. Kupfer-Bibel, In welcher Die Physica Sacra, Oder Geheiligte Natur-Wissenschafft Derer In Heil. Schrifft vorkommenden Natürlichen Sachen, Deutlich erklärt und bewährt . . . Augsburg and Ulm, 1731-1735. Schmitz, Emil-Heinz. Das Mikroskop (= Handbuch zur Geschichte der Optik, Ergänzungsbd. II). Bonn: Wayenborgh, 1989. Speck, Josef, ed. Grundprobleme der großen Philosophen. Philosophie des Altertums und des Mittelalters. 5th ed. Göttingen: Vandenhoeck & Ruprecht, 2001. Tozer, E.T. “Discovery of an Ammonoid Specimen describes by Robert Hooke.” Notes and Records of the Royal Society of London 44.1 (1990): 3-12. Turner, Gerard L’E. Essays on the History of the Microscope. Oxford: Senecio Publications, 1980. Turner, Gerard L’E. Mikroskope. Munich: Callwey Verlag, 1981. Zahn, Johannes. Oculus artificialis teledioptricus. Herbipoli, 1685 and Nuremberg, 1702. Zuylen, J. van. “The Microscopes of Antoni van Leeuwenhoek.” Journal of Microscopy 121.3 (1981): 309-28.
JOCHEN HENNIG
The Instrument in the Image: Revealing and Concealing the Condition of the Probing Tip in Scanning Tunneling Microscopic Image Design
Scientific images frequently do not reveal their origins and modes of production any more. The instrument is thus no longer visible as the prerequisite of experimentally generated images, and the isolated images seem to match the ideal of observation without any assumptions. According to Bettina Heintz and Jörg Huber, “the assumption of instrumentally mediated objectivity is the result of a cultural attribution . . . which either conceals the fact that the image is dependent upon the equipment and measurement techniques used to create it, or interprets this dependency as unproblematic.”1 The following essay argues that this attribution was also valid in principle at the time of the introduction and the establishment of the scanning tunneling microscope in the 1980s, but was realized only within a process. Whereas in the earliest years of this imaging technique, the published images drew attention to the role the instrument played in their creation, later, over a period of several years, forms of representation and publication emerged that allowed the concealment of any reference to the process. However, since tunneling microscopic images simultaneously offer insights into the analyzed sample and provide information about specific components of the instrument that need constant monitoring and adjustment, an inherent part of experimentation consists in examining the images in order to assess the instrument. Studies in the history of science have shown how individual instruments are transformed from tools into objects of study by the reflexive character of experimentation and also how individual instruments have come to be used in areas other than those for which they were original1
Bettina Heintz and Jörg Huber: “Der verführerische Blick.” Mit dem Auge denken. Ed. idem. Zurich and Vienna: Springer, 2001. 19.
The Instrument in the Image
349
ly developed.2 The emphasis of this article, in contrast, is on the constant monitoring of the state of instrument components as part of the laboratory routine. Taking the scanning tunneling microscope as an example, it analyzes the discrepancy between, on the one hand, the use of the image during an experiment as a means to assess the instrument, and on the other, a simultaneous consensus to conceal any reference to the instrumental creation process in published images. The technique of creating images with the scanning tunneling microscope has attracted much attention since the beginning of the 1980s and its use has become widespread. The award of the Nobel Prize to IBM scientists Gerd Binnig and Heinrich Rohrer in 1986 can be seen as broad recognition on the part of the interested public and specialist circles. The instrument was praised for its clear and graphic depiction of individual atoms in real space, a strong contrast to the diffraction images with atomic resolution created by the transmission electron microscope or field ion microscope.3 However, it took much laborious effort to design images based on scanning tunneling measurements in such a way that they appeared to portray individual atoms. Tunneling microscopy is not based on optics and the use of lenses, but on the interaction of a tip with a scanned surface. When voltage is applied, an electric current of the magnitude of several nanoamperes flows between the tip and the surface. This current overcomes the vacuum, that is, the non-conductive space between the tip and the surface. From the perspective of classical physics, electrons are unable to pass through the potential barrier the vacuum represents. According to quantum mechanic concepts of probability, however, they are able to tunnel their way through the potential barrier with a certain degree of probability. For this reason, the current between the tip and the surface is called the ‘tunneling current’ and the instrument is called the ‘scanning tunneling 2
3
Matthias Dörries has shown, for example, how scales and diffraction gratings became objects of study. Cf. Matthias Dörries. “Balances, Spectroscopes, and the Reflexive Nature of Experiments.” Studies in the History and Philosophy of Science 25 (1994): 1-36. Richard Staley has described how the Michelson interferometer came to be used in new areas. Richard Staley. “Michelson’s Interferometer. Experiment or Instrument?” Instrument – Experiment. Historische Studien. Ed. Christoph Meinel. Berlin and Diepholz: Verlag für Geschichte der Naturwissenschaft und der Technik, 2000. 192-200. Cf. the press release of the Royal Swedish Academy of Science on the award of the Nobel Prize for Physics to Binnig and Rohrer in 1986 for the development of the scanning tunneling microscope. Nobel Foundation. Press Release: The 1986 Nobel Prize in Physics. Online at: http://www.nobel.se/physics/laureates/1986/press.html (August 2005).
350
Jochen Hennig
Fig. 1: Three-dimensional paper model, based on a scanning tunneling microscopic measurement of a silicon surface.
microscope.’ The current is dependent upon the distance between the tip and the surface: the closer the tip is to the surface, the greater the current. When the tip is moved horizontally over a surface, the current flow can change. A feedback mechanism then corrects the vertical position of the tip and re-establishes the originally determined value for the current flow. The resulting vertical movement of the tip while it scans a surface creates the signal that serves as a basis for the compilation of tunneling microscopic images. Although this research received little notice at first, in 1982 an image of a 7x7 silicon surface made with the scanning tunneling microscope attracted much attention, especially within the surface physics community. To create this image, Binnig copied an analogue printout that recorded the path of the tunneling microscope tip in the form of a line. He then cut several strips out of this copy and pasted them together in intervals of about one millimeter to build a three-dimensional model. In the photograph, the model (fig. 1) was illuminated in such a way that it appeared to throw a visible shadow. The illumination thus created the impression of dimension and a recognizable topography in which maxima and minima were clearly differentiated. Since the magnitudes were known, individual maxima could be assigned to the dangling-bonds of individual atoms. At the same time, because the signals of individual atoms can be measured, this meant that both on the ana-
The Instrument in the Image
351
lyzed silicon surface and on the tip of the instrument a single atom took part in the measurement. The signal of tunneling microscopic measurements is always created by the current between the tip and the surface. Tip and sample reproduce each other, determine the display resolution, and appear in the image. If a pattern in the dimension of individual atoms appears in the image, as it does in the analysis of the silicon surface, this strongly indicates that a single atom is involved in the measurement process on the tip as well. Thus, in order to achieve atomic resolution with this imaging technique, it was necessary to develop a correspondingly sensitive tip. In their first experiments, Binnig, Rohrer, and Christoph Gerber, a technician, turned to available, technically manufactured tips like the ones already in use in field emission microscopy. The tunneling microscopic images created in this way did not exhibit the degree of surface resolution they had hoped for. From their investigation of these images, Binnig and Rohrer came to the conclusion that the tips of the field ion microscopes bent during use in the tunneling microscope. Gerber, who was responsible for building the first tunneling microscope in the IBM labs, then produced a stiffer tip made of tungsten wire with a grinding machine.4 The images created with this tip showed atomic resolution. After analyzing of these images, they concluded that the single, lowest atom must be responsible for the physical process in a tip ground in this manner. There was as yet no complete, quantitative theory capable of explaining the tunneling current between the tip and the surface at the atomic level. The conclusion that individual maxima could be assigned to the location of individual atoms was only possible by interpreting the image according to known dimensions and by interpreting the alignment of the maxima according to pre-existing crystallographic knowledge. In subsequent studies of different surfaces with the tunneling microscope, the samples were first characterized using other methods of surface analysis such as Auger spectroscopy, in order to exclude the possibility that they could be contaminated. The samples were clearly defined and their crystal structure was known. The measurement results, however, which were expressed in the form of images, served less to define the physical characteristics of the sample.5 Instead, scanning tun4 5
Gerd Binnig and Heinrich Rohrer. “Scanning Tunneling Microscopy.” Helvetia Physica Acta 55 (1982): 729. A striking example of the status of tunneling microscopic measurement in the middle of the 1980s can be found in the publication of an analysis of graphite by Binnig, Rohrer, and their colleagues, in which the scanning tunneling microscopic
352
Jochen Hennig
Fig. 2: Scanning tunneling microscopic image with asymmetrical maxima.
neling microscopists used these experiments to test and familiarize themselves with different construction principles in the instrument and, in particular, to study the interaction between the tip and the surface. During this phase, Binnig, Rohrer, and their colleagues analyzed a graphite surface. In their comments on the corresponding published analogue line image, (fig. 2) they wrote, “The left-right symmetry (steeper upward slopes from right to left) is not an artifact of data acquisition, but should be attributed to an asymmetric tunnel tip.”6 This quote shows that the experience, competence and authority that they had accumulated in interpreting the results enabled them to choose between different possible causes for one effect, namely the asymmetry of individual maxima. The asymmetry of the tip became visible as a characteristic of the instrument. It was a part of the image and it was discussed in the article that accompanied the publication of the image. A further publication showed the study of absorbed oxygen on nickel as a line image (fig. 3a) and a grey-scale image (fig. 3b). In the grey-scale image, the vertical path of the tip is given in six different
6
analysis is described as confirming the “tacit assumptions” of many adsorption experiments. Cf. Gerd Binnig, et al. “Energy-Dependent State-Density Corrugation of a Graphite Surface as Seen by Scanning Tunneling Microscopy.” Europhysics Letters 1 (1986): 31-36. Ibid. 32.
The Instrument in the Image
353
Fig. 3 a
Fig. 3 b Two different designs of one scanning tunneling microscopic measurement: as a result of a defect in the tip, the structures are less clearly defined in the upper parts than in the lower parts.
354
Jochen Hennig
shades of grey, which are separated by black lines, with lighter shades corresponding to higher positions. In this image, there is a line indicated by two arrows. The structures above this line are not as clearly defined as those beneath it. This can also be seen in the line graph. From studying these images, the authors concluded that the tip had been rammed into the sample at the position marked with the arrow during its scan of the surface of the sample. Instead of choosing to publish only the lower part of the image, which was scanned first, they published the entire image, drawing attention to the influence that the change in the tip had on the results. Thus the image explicitly directs attention to the condition of the tip and the sensitivity of the measurement. Despite early difficulties in reproducing Binnig and Rohrer’s results, around the same time as the measurements were made in the mid 1980s the scanning tunneling microscope began to be used in an increasing number of industrial and university laboratories. Two different approaches to the manufacture of tips were taken. Firstly, diverse and often individual manufacturing methods, such as Gerber’s method of grinding the tip were developed, or methods were adopted from field emission microscopy. Tips were also etched, broken, and ground or cut from wire. The scanning tunneling microscopic images served as a point of reference for an indirect assessment of these methods. Simultaneously, many scientists also developed particular skills. These did not become the subject of systematic publication, certainly due in part to the difficulty of verbally describing skills that had been acquired through practice.7 However, tip manufacturing techniques were also considered to be a secret of success that should not and need not be revealed, since the tunneling microscopic images served as proof of tip quality. One of the few articles actually published on this topic was entitled “The Art and Science and Other Aspects of Making Sharp Tips.”8 The title is unusual in that, although commonalities between art and science have been discussed in contemporary science studies,9 there is no place 7
8 9
Harry Collins summarized this point: “The major point is that the transmission of skills is not done through the medium of the written word.” Harry Collins. “Tacit Knowledge and Scientific Networks.” Science in Context. Readings in the Sociology of Science. Ed. Barry Barnes and David Edge. Milton Keynes: Open University Press, 1982. 54. Allan Melmed. “The Art and Science and Other Aspects of Making Sharp Tips.” Journal of Vacuum Science and Technology B 9.2 (1991): 601-08. The title of this volume also encourages one to view science and art together. For a critical analysis of the state of research on images in science and art, cf. Peter Geimer. “Weniger Schönheit. Mehr Unordnung. Eine Zwischenbemerkung zu ‘Wissenschaft und Kunst.’” Neue Rundschau 114.3 (2003): 26-38.
The Instrument in the Image
355
in specialist science publications for aspects of “the art of craftsmanship” in its literal meaning. In addition to these explorations into methods of tip-making, theorybased methods of studying and manufacturing tips were also developed.10 Thus, the IBM scientist Hans-Werner Fink used a field ion microscope to produce a tip with one defined atom at the bottom. He used the field ion microscope both as a tool for the manufacture of the tips, as well as for visual verification of their atomic structure. Despite fulfilling the requirement of creating a tip with a single atom at its bottom, Fink had no success when using these tips in the scanning tunneling microscope. He was unable to represent a graphite surface with atomic resolution, a standard test for checking and demonstrating the resolution capacity of an apparatus since the mid 1980s.11 This could be done, though, with tips that had been manufactured according to conventional and less elaborate methods, for example when they were etched in a beaker. Fink’s colleague at IBM, Erich Stoll, took the opportunity to write an article with the slightly derisive title, “Why Do ‘Dirty’ Tips Produce Higher-Resolution Images when Graphite is Scanned in a Scanning Tunneling Microscope?” Stoll explained that the graphite layers that formed on etched tips as contaminations possessed precisely the electronic characteristics needed for a high-resolution measurement of graphite surfaces. These contaminations were absent in Fink’s highly defined, vacuum-manufactured tips.12 Thus, in summary, an approach that tested different tips and drew conclusions about their suitability on the basis of scanning tunneling microscopic images proved to be more successful than an approach guided by theory. Although the latter fulfilled the necessary requirement of having a single atom at the tip, this alone proved insufficient since the complexity of physical processes on the tip could not be fully comprehended or predicted.
10
11
12
The concepts of explorative and theory-based experimenting used here are based on the discussion in Friedrich Steinle. “Exploratives vs. theoriebestimmtes Experimentieren. Amperes erste Arbeiten zum Elektromagnetismus.” Michael Heidelberger and Friedrich Steinle, eds. Experimental Essays – Versuche zum Experiment. Baden-Baden: Nomos, 1988. 272-97. Cf. Cyrus Mody. “Instruments in Training. The Growth of American Probe Microscopy in the 1980s.” Pedagogy and the Practice of Science. Producing Physical Scientists, 1800-2000. Ed. David Kaiser. Cambridge: MIT, 2005. 185-216. Erich Stoll. “Why Do ‘Dirty’ Tips Produce Higher-Resolution Images when Graphite is Scanned in a Scanning Tunneling Microscope?” Journal of Physics C – Solid State Physics 21.26 (1988): L921-L924.
356
Jochen Hennig
This recognition that the images provided a means to assess the quality and condition of the tip forms the background of the following discussion of how the design of published images evolved in the 1980s. In the early years of scanning tunneling microscopy the path of the tip was shown within the image. This rendered the production process of scanning tunneling measurements visible. During the 1980s, references to the experimental origins of the images became progressively less frequent in image design and rhetoric, until it seemed, instead, that the objects, the atoms themselves, were being portrayed. This led to a discrepancy between the images experimenters used in their laboratories to evaluate their measurements and thus also the quality of the tips and the images that were later published. This tendency can be traced back to the first measurements of the 7x7 silicon surface. Thomas Hartwig, who was employed at IBM as a pre-doc, worked on image processing largely independently of the people who were working directly with the scanning tunneling microscope. He scanned a line image from an earlier silicon analysis that had formed the basis for the paper model (fig. 1), and he converted the wavy lines into a top view in grey shades, in which the highest points were assigned the lightest shades of grey. The different grey tones were generated by the computer. Hartwig applied a technique that had been used in previous years in an IBM visualization laboratory that was located in the same IBM research center in Zurich as was the tunneling microscopy group. The technique had been developed independently of physical experiments. The graphic representation of grey tones on a monitor and in printouts was not an insignificant achievement at the time, but in 1980 the visualization group could not have known that in two years’ time it would be in demand from the Devices Department of the laboratory in connection with the newly developed tunneling microscope. The aim of this image processing was to create a height profile for the entire area, whereby the origin, the line-by-line scan with the tip, disappeared from the image. In unpublished printouts, one sees the transition from discrete lines, which are recognizable in the early stage (fig. 4a), to an area representation (fig. 4b) through the use of extrapolations.13 Erich Stoll not only thought about the processes between the lowest atoms on the microscopic tip and the surface, but also worked intensively on digital image processing. In an article published in 1985, he 13
I thank Hartwig Thomas for his willingness to discuss this with me (September 19, 2003) and for allowing me access to the original printouts in his private possession.
The Instrument in the Image
357
Fig. 4 a
Fig. 4 b Transition from a line image to an area representation, realized by computer-based image processing.
presented steps in image processing that could possibly be taken following an experiment with the tunneling microscope.14 In this article Stoll revealed that the images from the nickel analysis carried out by Binnig, Rohrer, and their colleagues (cf. fig. 3a, 3b) had already undergone several steps of alteration before being published. Stoll subsequently published the originally recorded printout and marked the spots that he had changed himself (fig. 5): contact between the tip and the surface (“1” in the illustration); the ascent of the individual scans (2); increasing distance between scans (3), low frequency vibrations (4), offset of consecutive scans (5), and high frequency noise (6). Through Stoll’s disclosure of this “prehistory,” one can situate the originally published images (fig. 3a, 3b) in a transitional phase. On the one hand, several aspects were deemed to be worthy of modification and were altered for publication. On the other hand, by the inclusion of both high and low resolution sections, these prepared, published ima14
Erich Stoll. “Information- and Image-processing of Scanning-tunneling-microscope Data.” SPIE 599 (1985): 442-50.
358
Jochen Hennig
Fig. 5: Original image of a scanning tunneling microscopic measurement; the marked spots have been changed for publication (fig. 3a).
ges do still show that the images were dependent upon the condition of the tip. The IBM logo that was created out of single atoms in 1989 (fig. 6) belongs to this transitional phase as well. This series of images demonstrates that the tunneling microscope could be used to move individual atoms.15 It was compiled in IBM’s Almaden laboratory in California using a tunneling microscope with the highest resolution possible at the time. In addition, it was capable of operating at a temperature of 4 Kelvin, that is, minus 269 degrees Celsius, and in the most perfect vacuum that could be created at the time. The published series of images was given a form of image design that appealed to macroscopic visual conventions, for example, by evoking similarity with an image of illuminated shells that threw shadows. This impression was created by coding differentials for the direction of tip movement, not the height, on a grey scale (cf. fig. 3b, 4a, 4b). Light grey tones marked the raising of the tip, 15
This demonstration of the scanning tunneling microscope moving single atoms received much attention, and the image formed the basis of nanotechnological utopias that predicted the construction of machines out of individual atoms.
The Instrument in the Image
359
Fig. 6: Image sequence for the production of an atomic IBM logo with a scanning tunneling microscope; in the images c, d, and e impurities on the tip caused the apparent double mapping of each atom.
medium grey was used to depict movements that were nearly horizontal, and dark grey tones were used to mark the sinking of the tip. In images c, d, and e of the series showing the construction of the logo (fig. 6), every atom seems to be reproduced twice. This is due to the contamination of the tip during this daylong experiment. The interference was not mentioned in the text of the publication in 1990, in contrast to articles from the mid-1980s, which explicitly discussed the effect of changes in the tip. Still, for the compilation of the last image of the series, which was to become an icon of nanotechnology,16 it proved possible to clean the tip. This was done by briefly increasing the voltage between the tip and the 16
For a discussion of this image, cf. Jochen Hennig. “Vom Experiment zur Utopie. Bilder in der Nanotechnologie.” Instrumente des Sehens (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, Bd. 2,2). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2004. 9-18.
360
Jochen Hennig
surface, which caused the contamination to fall off the tip.17 Cleansing the tip in the course of an experiment required the development of particular skills, much as tip production had. It consisted of briefly rotating a potentiometer to create a momentary increase in voltage. Scientists and technicians were able to develop this skill and the corresponding sense for the instrument because they could read the success of their strategies for cleaning the tip in the images and used this knowledge to compile images like the final one in the series showing the creation of the logo. Without a single reference to the instrument and the interaction between the tip and the surface, individual atoms themselves seem to appear as spheres. As the tunneling microscope became established in the 1990s, images that openly showed the influence of the condition of the tip disappeared from publications. However, as the use of flawed or contaminated tips remained an unavoidable part of laboratory work, experimenters developed routines for dealing with this circumstance. Whereas in the early years of tunneling microscopy, samples of easyto-scan and well-known materials such as the graphite surface were used to check the tips,18 in the 1990s defined, commercial test samples became available as well. If during an experiment there is a question as to whether a particular, unexpected aspect of the image should be attributed to the sample or the tip, the same tip can be used in a subsequent measurement to scan the test sample, which has a known structure. By comparing this test image with other images from the standard sample, the experimenter can draw conclusions about the condition of the tip.19 While it continues to be ‘problematic’ for tunneling microscopists to interpret whether aspects of the image are the result of defects in the tip, the published images seem to suggest, in agreement with the observation by Heintz and Huber stated at the beginning of this essay, that this dependency is ‘unproblematic.’ Translation: Elizabeth Neswald 17
18
19
Erhard Schweizer, who participated in this experiment, communicated these experimental details to me in an interview on 8 October 2003. They were not mentioned in the publications. Sang-il Park, Jun Nogami, and Cal F. Quate. “Effect of Tip Morphology on Images obtained by Scanning Tunneling Microscopy.” Physical Review B 36.5 (1987): 2863-66. Reinhard Guckenberger. “Test Structures and Calibration.” Procedures in Scanning Probe Microscopies. Ed. Richard J. Colton et al. Chichester: Wiley, 1998. 9-21.
The Instrument in the Image
361
WORKS CITED Binnig, Gerd and Heinrich Rohrer. “Scanning Tunneling Microscopy.” Helvetia Physica Acta 55 (1982): 726-35. Binnig, Gerd et al. “Energy-Dependent State-Density Corrugation of a Graphite Surface as Seen by Scanning Tunneling Microscopy.” Europhysics Letters 1 (1986): 31-36. Collins, Harry. “Tacit Knowledge and Scientific Networks.” Science in Context. Readings in the Sociology of Science. Ed. Barry Barnes and David Edge. Milton Keynes: Open University Press, 1982. 44-64. Dörries, Matthias. “Balances, Spectroscopes, and the Reflexive Nature of Experiments.” Studies in the History and Philosophy of Science 25 (1994): 1-36. Fink, Hans-Werner. “Mono-atomic Tips for Scanning Tunneling Microscopy.” IBM Journal of Research and Development 30.5 (1986): 460-65. Geimer, Peter. “Weniger Schönheit. Mehr Unordnung. Eine Zwischenbemerkung zu ‘Wissenschaft und Kunst.’” Neue Rundschau 114.3 (2003): 26-38. Guckenberger, Reinhard. “Test Structures and Calibration.” Procedures in Scanning Probe Microscopies. Ed. Richard J. Colton et al. Chichester: Wiley, 1998. 9-21. Heintz, Bettina and Jörg Huber. “Der verführerische Blick.” Mit dem Auge denken. Ed. idem. Zurich and Vienna: Springer, 2001. Hennig, Jochen. “Vom Experiment zur Utopie. Bilder in der Nanotechnologie.” Instrumente des Sehens (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, Bd. 2,2). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2004. 9-18. Melmed, Allan. “The Art and Science and Other Aspects of Making Sharp Tips.” Journal of Vacuum Science and Technology B 9.2 (1991): 601-08. Mody, Cyrus. “Instruments in Training. The Growth of American Probe Microscopy in the 1980s.” Pedagogy and the Practice of Science. Producing Physical Scientists, 1800-2000. Ed. David Kaiser. Cambridge: MIT, 2005. 185-216. The Nobel Foundation. Press Release: The 1986 Nobel Prize in Physics. Online at: http://www.nobel.se/physics/laureates/1986/press.html (August 2005). Park, Sang-il, Jun Nogami, and Cal F. Quate. “Effect of Tip Morphology on Images Obtained by Scanning Tunneling Microscopy.” Physical Review B 36.5 (1987): 2863-66. Staley, Richard. “Michelson’s Interferometer. Experiment or Instrument?” Instrument – Experiment. Historische Studien. Ed. Christoph Meinel. Berlin and Diepholz: Verlag für Geschichte der Naturwissenschaft und der Technik, 2000. 192-200. Steinle, Friedrich. “Exploratives vs. theoriebestimmtes Experimentieren: Amperes erste Arbeiten zum Elektromagnetismus.” Experimental Essays – Versuche zum Experiment. Ed. Michael Heidelberger and Friedrich Steinle. Baden-Baden: Nomos, 1988. 272-97. Stoll, Erich. “Information- and Image-processing of Scanning-tunneling-microscope Data.” SPIE 599 (1985): 442-450. Stoll, Erich. “Why do ‘Dirty’ Tips Produce Higher-Resolution Images when Graphite is Scanned in a Scannning Tunneling Microscope?” Journal of Physics C – Solid State Physics 21.26 (1988): L921-L924.
BRUNO BACHIMONT
Formal Signs and Numerical Computation: Between Intuitionism and Formalism. Critique of Computational Reason 1. Intuitionism and Formalism The Legacy of the Seventeenth Century Signs constitute valuable supports for reasoning, particularly logical and scientific reasoning. It was in the seventeenth century that the principal conceptions of the sign and its role in reasoning were forged. Here, one can distinguish between the geometric intuitionism of Descartes, the arithmetic and algebraic formalism1 of Leibniz, and thirdly philosophical reasoning, which, as Immanuel Kant will emphasize later, as a worthy heir to these reflections on the sign, constitutes an intermediary position.2 Cartesian geometric intuitionism corresponds to the fact that the sign presents, directly and immediately (i.e. without mediation), the content of which it is the sign. This is the case, for example, with the geometric figure. Reasoning is carried out guided by the figure, in so far as it shows directly that of which it is the case. By following the symbolic letter of the figure, the geometer has no need to mobilize signified and absent entities; the signification is here immanent in the figure that shows what it signifies. For this reason it enables reasoning to be followed in an assured and veracious manner, where the power of the form is at the service of a defined content, constantly adherent to the signifying form, and directly apprehensible. The exactitude of the reasoning, and the foundation of its truth, consequently rely on evidence (etymologically, what comes into view is thrown into relief). The form of the figure only has value in so far as it makes visible. 1 2
Cf. Yvon Belaval. Leibniz critique de Descartes. Paris: Gallimard, 1960. Cf. Franck Piérobon. Kant et les mathématiques. Paris: Vrin, 2003.
Formal Signs and Numerical Computation
363
The algebraic formalism of Leibniz follows a different track. Indeed, with the evidence remaining sullied by arbitrariness and subjectivity, the formalist mathematician mobilizes a symbolism that he handles using a combinatorics regulated by formal laws, i.e. applied uniquely by virtue of form, independently of the content signified. This recourse to symbolism gives rise to a cognitive economy making it possible to carry out complex reasoning, where not all the elements can be present to the intellect: All human reasoning is accomplished by means of certain signs or characters. It is not only the things themselves, but also the ideas of the things that the intellect cannot, and should not, always observe in a distinct way; this is why one places signs in their place, in order to abbreviate . . . Therefore, names have been given to contracts, figures, to the various kinds of things, as well as signs to numbers in arithmetic, and to sizes in algebra, so that if experience and reasoning one day allows us to discover certain things, one can consequently combine in all confidence the signs of one with the signs of others.3
Reasoning is carried out with confidence, not because it refers to a content, manifested and rendered visible by a geometric symbolism, but because it mobilizes the handling of effective material signs, which can be sufficiently considered in themselves, and independently of what they signify. We must then realize that the tests or experiences carried out in mathematics to guard against mistakes in reasoning . . . these tests are not made on a thing itself, but on the characters which we have substituted in place of the thing.4
This is what already happens in arithmetic, though in fact above all in algebra. Departing from a correspondence initially established between symbols and entities they define (external to symbolism), the algebraic operation is carried out without consideration of these entities in order, eventually, to reconnect the corresponding entities with the symbols by which the result is reduced, and to find the solution sought for.5 In a certain way, intuitionism and formalism come together. In both cases, 3
4
5
Gottfried Wilhelm Leibniz. “Projet de Préface à la science générale.” Die philosophischen Schriften von Gottfried Wilhelm Leibniz. Vol. 7. Ed. Carl Immanuel Gerhardt. Hildesheim: Olms, 1890. 204. Gottfried Wilhelm Leibniz. Opuscules et fragments inédits de Leibniz. Extraits des manuscrits de la Bibliothèque Royale de Hanovre. Ed. Louis Couturat. Paris: F. Alcan, 1903. 154. Kant paraphrases this quite succinctly. Cf. Immanuel Kant. “Untersuchung über die Deutlichkeit der Grundsätze der natürlichen Theologie und der Moral.” Moses Mendelssohn. Abhandlung über die Evidenz in metaphysischen Wissenschaften, welche den von d. Kgl. Acad. d. Wiss. in Berlin auf d. Jahr 1763 ausgesetzten Preis erhalten hat. Berlin, 1764. § 2, 73.
364
Bruno Bachimont
reasoning in relation to content is only carried out by faith in what one sees and deals with tangibly. From the intuitionist point of view, the content is also there in front of one’s eyes, adherent and immanent in symbolism; from the formal point of view, the content has no need of consideration. In short, whether the content is present or one can do without it, one only reasons due to the presence of symbols. It is another matter in the case of philosophical reasoning. Indeed the latter, depending on natural language, has to mobilize concepts by means of words, and the intellect continuously has to be careful to keep the content, signified but not manifested by the symbols, in view. In contrast to intuitionism, the words and symbols of conceptual and philosophical reasoning do not directly manifest the content. In contrast to formalism, one cannot avoid considering it, in order to carry out reasoning. This type of reasoning does not seem to have been able to find a method that would guide it with assurance and certainty on the path to truth, since it must constantly protect itself from verbalism, that is to say, the formal tendency to trust words and their combination without verifying the implications at the level of content, and mysticism, i.e. the intuitionist tendency to depend on a content that one can never see, communicate, or share. Caught between a form imperfectly reflecting the content, and a content which is not directly accessible, the discursive reasoning based on the linguistic sign benefits from a combinatorics whose blindness to content leads to a blindness of reasoning. That is why formal efficacy can only be a support for the exercise of thought and its algebraic substitution. An Analogous Contemporary Situation: Calculation and Signification Computer science can be understood as an automation of reasoning, carried out in formal systems6 – algebraic systems that formalize logical and mathematical reasoning. These systems algebrize mathematical activity itself, not the objects which the latter normally deals with. From this point of view, computer science is a direct and extreme consequence of Leibniz’s algebraic formalism, which, as we know, was taken up again, during the profound crisis of mathematics, by David Hilbert: the blind use of signs considered independently of their content allows calculations to be carried out.7 Therefore, formalism definitively won the match against Cartesian intuitionism, and as a result, became the only 6 7
Cf. John Haugeland, ed. Mind Design. Cambridge: MIT Press, 1981. David Hilbert. “Über das Unendliche.” Mathematische Annalen 95 (1926): 161-90.
Formal Signs and Numerical Computation
365
adequate conception for understanding the status of computer programs and their operations. However, it is not like this. In fact, we need to distinguish between two very different ways of using computational systems. In the first, one observes a quantity of mathematical entities presented by a precisely defined symbolism. This symbolism, governed by formal rules (independently of signified content), can be put to work via a program. The blind behavior of the program remains nonetheless adequate to the content, in the manner of mathematical or Leibnizian formalism. The result of the program is intelligible since one can assign, with exactitude and precision, an entity in the observed mathematical world to each sign of computational-formalism. In the second situation, one does not use symbols associated with a mathematical model, but freely interprets symbols according to rules borrowed from general language or the specialized language of a field of activity. These symbols, belonging to the vocabulary of these languages, have no precise signification (in the sense of mathematical and logical formalism), and it is not possible to attribute to them, in any univocal way, a signified and associated entity. In other words, although one has formalism at one’s disposal in as far as the symbols are used uniquely in relation to their form, and independently of content, these symbols, in contrast with algebraic symbolism, have no precise signification. In the latter situation, one rediscovers difficulties analogous to those of philosophical reasoning: since formalism is not systematically linked with content, one either risks falling into verbalism by taking the symbols appearing on the computer too literally, or into mysticism by believing that the system is targeted at the same entities as those we connect with the symbols used. This situation corresponds, today, to the majority of the circumstances of computational tool use, whether in logistics, administration, or documentation, etc. Formalisms are mobilized whose signification is not systematically established, so that it is impossible to establish with rigor and certainty what the machine is trying to say when it suggests a result, nor to know what is signified. In other words, outside scientific computation, one finds oneself once again faced with results whose intelligibility is not predetermined. In this case, one can produce non-sense but also new and unanticipated symbolic configurations whose interpretation enables the emergence of new thoughts and conceptions, just as one can understand sentences one has never said or heard. It is precisely these kinds of situations we want to treat in this article. How should we understand the nature of what we do with our com-
366
Bruno Bachimont
puters if it is a case not only of automating our algebraic symbolisms derived from mathematical sciences, but of mechanizing the use of symbols as if they were algebraic (i.e. by disregarding their signification during their use), when they are not (to understand them is to observe them in their physical materiality together with their meaning without ever separating these two things). This objective is important inasmuch as it is necessary to supplement the invention that makes algebraic combinatorics possible (as Leibniz suggested; see above) with one that makes this pseudo-algebraic combinatorics possible. In order to discuss this invention, we suggest considering these formalisms as writings and intellectual tools, i.e. as symbols whose use can be purely formal if one stays with their material and physical nature, but whose interpretation cannot be related to algebraic rules. It is necessary, therefore, to understand how these computational or numerical tools put the combinatorics of symbols at the service of the power of the invention of sense. 2. Numerical Technologies and Writing In the long history of the intellectual tools that help us to think, computer science and the numerical occupy an important place, whose implications still need investigating. Indeed, if the scientific and technical nature of the numerical and of calculation rests on firm ground, the consequences resulting from its extensive use in the most diverse sectors of human activity are still difficult to measure. Numerical technologies inscribe themselves in the movement of the externalization and prostheticizing of thought8 so that intellectual operations can be consigned and confined to those material tools and instruments, thus unburdening thought and enabling it to turn its attention to other things. However, by being confined to material instruments and supports, intellectual tasks change their nature, and the intellect, when reappropriating the result, finds something different from what it would have found had it taken on these tasks itself. For example, if, instead of memorizing a talk, one writes it down, thus entrusting it to the material support of paper, then the reading of the resulting text allows the retrieval of other properties than the proc8
Cf. Bernard Stiegler. Technics and Time, 1: The Fault of Epimethus. Trans. Richard Beardsworth and George Collins. Stanford: Stanford University Press, 1998; André Leroi-Gourhan. Gesture and Speech. Trans. Anna Bostock. Cambridge: MIT Press, 1993.
Formal Signs and Numerical Computation
367
ess of recall. Thanks to techniques of writing and editing, reading has access to the exact content, while the memory can remain fallible. The written text, furthermore, enables us to see, together and arranged in the same space, sequences of words that would have been separated by the succession of speech. The text offers simultaneous access to a content, reassembled in the synoptic space of writing; content which, up to this point, remained dispersed in the temporal linearity of speech. By spatializing speech, writing de-linearizes it, and enables the observation of the content in the two-dimensional spatiality of the page. Nevertheless, the written loses the intonations, prosody, and emotional markers, of which the memory can retain a more or less true recollection. If the written increases the intelligibility of speech by making it visible, then it also removes levels of comprehension. The written gives us the material to think differently, to think something else, and to bring about new intellectual tasks and objects. Consequently, the idea starts to crystallize that our intellectual tools, according to their nature and properties, help us to think differently, just as mechanical tools allow us to produce different material objects. And, just as there is a history of techniques and the objects produced by them, there is also a history of our intellectual tools and their corresponding modes of thought. If, like writing, the numerical marks a rupture in the history of the tools of thought, then the question arises of how we think differently when we have numerical tools at our disposal, how we produce new intellectual objects, and how we elaborate concepts which would remain inconceivable without such a numerical mediation. However, one should not assume that the technological mutation of intellectual tools necessarily leads to a supplementation, to an extension of our cognitive field. It could also result in a deficiency of intelligibility, in a loss of sense, a disorientation. The possibilities opened up by a technological mutation, before being actualized, could result in a lack. What lack is introduced by the numerical? It seems that the supplementation produced by the numerical can doubtlessly be better understood in relation to the lack we come across when we discuss the numerical instrumentalization of our way of thinking. This lack is the mark of a rupture vis-à-vis prior practices, and describes within itself the new horizons that we might attempt to explore. And finally, as the last condition: numerical mutation is neither a progression nor a regression, but a mutation that imposes itself on us even if we are its author. The numerical is neither better nor worse than writing, for example, but surely just as important. Therefore, it is not
368
Bruno Bachimont
appropriate to lament a possible regression, or to extol a hypothetical progression, but rather to endow ourselves with concepts that allow us to think the numerical. 3. From Graphic Reason to Computational Reason It might seem surprising to connect the term “reason” with a technical principle: writing in one case, calculation in another. Thinking and particularly rationality is usually considered an activity belonging to the intellect, rightfully or factually independent of its cultural or historical environment. In other words, the thinkable does not depend on the technical environment of the intellect, even if, in a contingent way, what is effectively thought can be affected by the material conditions of reason. Consequently, to evoke, as Jack Goody9 proposes, a graphic reason, and, as we suggest, a computational reason,10 is paradoxical, since one would make an activity that rightfully only depends on itself dependent on a technical principle. Nevertheless, work on writing has made it possible to show that the technical innovations that have marked its history have had direct consequences on the thinkable (and not only on the thought). The consequences linked to mastering the technique of writing can be observed in two aspects: anthropological or historical. Jack Goody is, without doubt, the anthropologist who has contributed the most to our understanding of the role of writing in the emergence of certain cognitive operations or ways of thinking specific to writing. In his famous work, for which the French translator chose the title Raison graphique (i.e. graphic reason; the original title being The Domestication of the Savage Mind), Goody shows to what extent cultures endowed with writing reason differently from so-called oral cultures. Instead of attributing to these differences an axis of development and continual progress, where the oral cultures would occupy a lower position than the writing cultures (the former embodying a state prior to the latter), Goody shows that it is rather a case of differences proceeding from distinct technical instrumentalizations. 9
10
Cf. Jack Goody. The Domestication of the Savage Mind. Cambridge: Cambridge University Press, 1977; idem. The Logic of Writing and the Organization of Society. Cambridge: Cambridge University Press, 1986. Cf. Bruno Bachimont. “L’intelligence artificielle comme écriture dynamique. De la raison graphique à la raison computationnelle.” Au nom du sens. Ed. Paolo Fabbri and Jean Petitot. Paris: Grasset, 1999. 290-319.
Formal Signs and Numerical Computation
369
To what extent does writing bring about particular cognitive operations, a graphic rationality? Essentially, writing brings to content a spatial synopsis, enabling the recognition of relations and properties that remain untraceable in the linear succession of the temporality of speech: writing makes relations visible that are not perceivable when listening to speech. Indeed, by bringing a spatial two-dimensionality to the representation of content, the intellect can simultaneously access different parts of the content, independently of the order joining these parts in the oral flux. Consequently, what is dispersed in time becomes contiguous in space, the eye being able to navigate freely and recognize identities among the elements of the content (e.g. words possessing the same root but different inflexions). While a sentence containing the words “rosa, rosae, rosam, rosas, etc.” in linear succession is very unlikely, as much as orally it remains very difficult to recognize that the different words belong to the same declension where one can, moreover, expose the structure (the different inflexions), the written representation allows the de-linearization of the discourse and the isolation of units, which one can then compare and juxtapose in the intellect. Writing and Synthesis Writing enables spatial synthesis, i.e. the process of putting (-thesis) together (syn-) units otherwise dispersed, and combining them in the same structure or category. From a certain point of view, writing makes possible what Kant, in the Critique of Pure Reason, describes as the three syntheses, enabling consciousness to combine individual sensations or representations in the unity of a concept. As is known, Kant distinguishes the synthesis of apprehension in intuition, the synthesis of reproduction in the imagination, and the synthesis of recognition in the concept. These syntheses are necessary, since the intellect sees itself given in experience a spatial-temporal diversity, i.e. a content dispersed in time and space. This dispersion prevents something being seen in this diversity, not even a something. It lacks the unity allowing something to be recognized. To overcome this dispersion, it is necessary to reassemble the elements that compose it, to apprehend them globally, unitarily. This will be the task of the synthesis of apprehension in the intuition, i.e. to put together (this is the synthesis), in the same apprehension, that which is given in the intuition. But for this to be possible, it is necessary to put together, in the same unity, what is dispersed in time. The dispersion in time implies, however, that the diversity passes, the present
370
Bruno Bachimont
continuously sinks into the past. If one wants to carry out a synthesis of a temporal diversity, it is necessary to maintain in the present what is drifting into the past, so that all elements of the diversity are given together, and are simultaneously available for one and the same apprehension. Maintaining the past in the present occurs by reproducing it, thanks to the power of imagination, in the present. The imagination repeats the past in the present to make it available to the present, and graspable as a whole. But this repetition cannot occur randomly; one can only repeat, from the diversity sinking into the past, what is coherent and relevant to the constitution of a unity. That is the task of the concept: to provide a principle of selection, having the insight of what it is necessary to reproduce in the imagination. So the concept is a rule giving the principle of a reproduction in the imagination. Should we undertake a phenomenological description of this process, and therefore depart from the letter of Kant’s thought, one can reformulate thus: I look at, for example, my book on the table and go around the table. My perceptions continuously replace each other and sink into the past. My imagination repeats what is given to my perception, but not the whole scene, only what my attention retains, i.e. the book, such as my concept of the book allows it to be characterized. This concept of the book allows me to select from the scene perceived what should be repeated in the imagination, and to put these together with the other views of the book as they are given in the course of all perceptions. These views then give me the perception of the book as a spatial three-dimensional object with a cover, pages, etc. Writing makes possible, technically, the realization of what Kant’s three syntheses carry out. By spatializing speech, writing maintains in the present the elements that compose it. By means of symbolic (e.g. alphabetic) transcription, it selects what is given in the sound-perception, although only retaining the phonemes, e.g. independent of the prosody. Thus, in this particular case of phonetic writing, the phoneme is the concept allowing the transcription that secures, in the space of writing, the permanence of the phonetic given. At this point, another perception is possible: to continue the example from above, thanks to the concept of “the same lexical form,” I can combine in the space of my page the words dispersed in the transcription, but accessible simultaneously, showing the same structure, for example, “rosa” and “rosae”, enabling me to expose a paradigm of declension. In other words, writing not only enables the intellect to accomplish what Kant describes in his three syntheses, but also to constitute, in the phenomenological sense of the term, new concepts.
Formal Signs and Numerical Computation
371
Writing is a technique that exposes the intellect to new synthetic configurations making possible the constitution of new concepts. This inverts the Kantian order. According to Kant, in order to perceive something, I must apprehend the intuition globally, therefore I must reproduce in the imagination, and therefore I must mobilize a concept. Doing this, the concept (including all other a priori structures, particularly space and time) is the condition of synthetic apprehension, and not the result. But if the synthesis comes about through technical efficacy, then it is not conditioned by the concept, at least not the concept for which it enables the constitution. Indeed, if writing corresponds to the concept of phonetic transcription, it does not correspond to the concept of grammar or declension, for which it nevertheless enables the constitution. Writing, building on the execution of an intention, and of certain conceptuality, enables the constitution and elaboration of others. The concept is, therefore, both condition and result of the technical synthesis, keeping in mind that technology proposes new synthetic configurations to the apprehension of the intellect. Technology and Synthesis Now we can try to transfer to technology what we have just observed in relation to writing. Technology allows, through the structuring that it brings to the time and space of our experience, the constitution of new knowledge and new concepts. Far from only being a simple application of theories or concepts elaborated independently of it, technology is the condition for the elaboration of knowledge. By instrumentalizing our experience by means of repeatable methods and tools that extend our action, technology transforms our relation to the world, and leads us to think it differently, to the extent that we do not only think differently a world that stays the same, but that we constitute new worlds, with large or small ruptures between. The Structures of Graphic Reason This is why Jack Goody insists that writing induces a particular mode of thought, and a specific relation to the world. According to Goody, writing produces three principle types of conceptual structures that condition our modes of thought. These are the list, the table, and the formula. The list enables a de-linearization of discourse, isolating units that one sub-
372
Bruno Bachimont
sequently arranges in an enumeration. The list enables the gathering into the same unity of what is dispersed in discourse, inducing a classification and categorization. To make lists is to retain one element among others, noting that these elements have something in common: they belong to the same class, the same category. By favoring the structure of the list, writing induces a relationship to the world that arises from classifying reason: to think the world is to organize it into classes and hierarchies, that is, to order and arrange it. The world of writing is the cosmos of antiquity, like the universe (in the sense of totality; universum referring to the whole of the things considered globally), organized, coherent, and harmonious. As is known, kosmos originally signified “ornament” and provided the root “cosmos” for “cosmology” as well as “cosmetic”.11 Could this harmonious universe be an artifact of alphabetical writing? This is a hypothesis provoked by the classificatory possibilities induced by lists, whose constitution is made possible by writing. The table is the act of representing a totality of relations between unities by their respective position within the two-dimensional space of writing: to be left or right, above or below, are the two types of spatial relations which enable the semantic combination of units positioned in this way. In a table, the unit occupying a particular section takes on a predetermined signification, or at least one conditioned by the position of the section within the table. The mode of thinking induced by the table is therefore the system: a table specifies relations between the sections, and allows us, for instance, a priori and in a systematic way, to predict the value which a section must occupy from the position of this latter. The most famous example is probably Mendeleev’s periodic table, whose systematic nature made possible, at the time of its formulation, the discovery of future elements (such as uranium). Finally, the formula. The formula is a procedure making procedures of reasoning possible in relation to form alone without having to take account of signification. Since form includes in its structure that which should be retained from the observed significations, it is sufficient to use the form to carry out procedures of reasoning in relation to the content or signification. It is what lies at the foundation of formal logic and, more generally, mathematics. The problem consists less of knowing if formalism enables reasoning independent of content, the latter being itself questionable (does it really exist?), but of being able to depend on the form. 11
Rémi Brague. Wisdom of the World. The Human Experience of the Universe in Western Thought. Trans. Teresa L. Fagan. Chicago: University of Chicago Press, 2003.
Formal Signs and Numerical Computation
373
From the Graphic to Calculation Dependence on form is the attitude at the basis of all formalisms, notably those at the origins of computer science and the numerical. When, at the end of the nineteenth century, the crisis of the foundations of mathematics occurred, mainly caused by the paradoxes arising from set theory, numerous mathematicians sought the means to overcome this crisis. David Hilbert proposed considering mathematics in relation to its writing, and to search for procedures enabling the control of this writing. The objective was to assure that when a mathematical expression is produced, it is not possible to derive a contradictory expression. To achieve this, Hilbert considered mathematical writing in a purely formal way,12 that is, by only considering the signs used independently of their signification. But now the number of signs is limited, the expressions and the mathematical texts are limited, the time mobilized by the mathematicians is limited: one finds oneself again using a limited number of signs within a limited time; in short one makes sign-combinatorics. Therefore, it is sufficient to find purely formal rules which demand no mathematical inventiveness or particular comprehension to apply them, which allow us to make sure that the signs of a mathematical expression or text do not allow the deduction of contradictory expressions. In contemporary terms, one would say that it is necessary to find a program that allows us to prove that a mathematical expression does not entail a contradiction. This objective established by Hilbert could not be achieved. This “program” of Hilbert’s (program in the sense of objective or a work “program”) nevertheless enabled the elaboration of the idea of a system of signs (the expressions) operating in a purely formal way via formal rules. Alan Turing, the father of computer science, developed this line further by proposing a theoretical machine conceived in the following way:13 one is provided with an unlimited memory-band composed of sections containing only a single symbol (viz. the sheet of paper on which the mathematician works), a reading and writing head moves on the memory band from section to section (one at a time) being able to read and write one symbol (the pen of the mathematician), and finally the internal condition of the reading-head (the mental condition of the mathematician). A purely formal program enables the determination of 12 13
Ibid. Cf. Alan M. Turing. “On Computable Numbers, with an Application to the ‘Entscheidungsproblem.’” Proceedings of the London Mathematical Society 2.42 (1937): 23065 and its corrections in the following issue 2.43 (1937): 544-46.
374
Bruno Bachimont
what the reading-head can write, and how it must move in relation to what it reads, and considering its internal condition. In other words, in relation to what he sees and thinks, the mathematician writes a symbol on a piece of paper. The metaphor should not disconcert us. The manipulation of signs on the memory-band are purely formal, and the functioning of the Turing machine rests, not on interpretation or signification associated with symbols, but uniquely on their form. In other words, the formalism arising from the structure of the formula, and rendered possible, according to Goody, by writing, has brought about the idea of automatic systems which use formal signs: an automatic formal writing, which, as it were, writes by itself. A Computational Reason? This idea engendered computer science, a technique enabling the automatic manipulation of symbolic inscriptions, whether these represent numbers, letters, or anything else. To what extent does the recourse to calculated representations induce a particular rationality? We will approach this question from two angles. Firstly, what would be the cognitive or phenomenological use of formal calculation and computer science for the understanding, following the example of writing, which offers a spatial synoptic synthesis of what is dispersed in time?14 Secondly, what would be the fundamental thought structures called up by computer science, following the example of what the list, the table, and the formula are for writing? If writing enables the synthesis of time in space by gathering what is dispersed in time (flux of speech) into the unity of a spatial synoptic representation, thereby making it possible for the intellect to trace synthetic configurations and constitute new concepts, computer science, for its part, enables the deployment of space in time. In fact a program is nothing other than a dispositif regulating a course in time (calculation or the execution of a program), departing from a structure specified in space (algorithm or program). The algorithm specifies that once the initial conditions are fulfilled, the result cannot fail to be obtained, according to the given complexity. The program is therefore a means to secure the future, to eliminate the uncertain or improbable, to bring it back under control. The time of computer science is therefore not an availability of 14
Cf. Bruno Bachimont. Herméneutique matérielle et artéfacture. Des machines qui pensent aux machines qui donnent à penser. Ph.D. École Polytechnique, Paris, 1996.
Formal Signs and Numerical Computation
375
what is to come, just as improbable as that might be, but rather the negation of the future in its opening up, to reduce it to what can be obtained from the perspective of the present. Calculation is becoming – in the opening, the availability of being – reduced to what is to come in the certainty of formalized prediction. Calculation establishes a kind of equivalence or correspondence between time and space: time becomes the time that is necessary for the systematic exploration of a space of calculation, as the path through all possible cases of a combinatorics; space becomes the space which is traversed in a certain number of stages specified by calculation. But space and time are dual: space is what one can pass through by means of the stages of the calculation; time, the necessary stages for the traversing of space. Under these conditions, what is, corresponding to the synoptic spatiality of writing, the cognitive function of calculation? We propose the notion of a systematic exploration. The calculation is what enables the systematic traversal of a space of possibles. It is this notion of a systematic exploration that enables us to derive the conceptual structures characteristic of computational reason. We propose considering the notion of the program, the network, and the layer. The Structures of Computational Reason The program is to computational reason what the list is to graphic reason. As much as the list makes categorization and classification possible, offering a spatial synopsis, the program enables the determination of a systematic path. The execution of the program is therefore merely the temporal deployment of the symbolic spatial structure that is the program. The network is to computational reason what the table is for graphic reason. While the table proposes a structuring and systematicity among the contents, divided into the sections of the table, the network proposes a mode of communication and distribution among the sections of the table. It is a dynamic table. Finally, the layer is to computational reason what the formula is to graphic reason. The formula enables the consideration of form while bypassing content; the layer enables the consideration of calculative relations among units while bypassing the subjacently implied calculations. The notion of layer in computer science, via that of implemention and compilation, enables the representation of formal structures, bypassing the induced elementary calculations, just as the formula enables an abstraction from sense (fig. 1).
Bruno Bachimont
376 Graphic Reason
Computational Reason
List
Program
Table
Network
Formula
Layer
Fig. 1: Comparison of graphic and computational reason.
These cognitive structures are fundamental and influence henceforth our modes of thought. Graphic reason produced a classifying reason; computational reason produces thinking in networks, and the time of prediction. For graphic reasons, the network is not a structure of intelligibility: the network, which because of its complexity withdraws from spatial synopsis, is a labyrinth where one loses oneself. It is a figure of the irrational, and not a way to think the world. Interaction and communication, in accordance with the structure of the networks, have become intelligible since calculation enables the reduction of the complexity, and the traversal of the whole of the possibles, induced by the networks, by means of the programs that specify the behavior of the calculation. In the same way, the notion of layer is equally a way of reducing complexity, and bringing a nearly infinite number of formal calculations back to structures which are more intelligible to humans. The structures in a network and layer, via the programs which realize them and render them effective, enable the approach of the real, not as a structure, hierarchized and organized into classes, but as a dynamic force deploying a subjacent rationality and order. The world is merely the execution of programs, which temporalize the relations they specify. Not that it is necessary to imply that there is a single subjacent program, but, inversely, that many orders interact together. Since these interactions are not necessarily predictable or coherent, it is necessary to search for the program and reconduct the search for a calculated order. If the taxonomy of species can be an illustration of the thought induced by graphic reason, then the genetic code is that of the thought induced by computational reason.
Formal Signs and Numerical Computation
377
4. Critique of Computational Reason The Problem of Intelligibility The principle virtue of calculation consists of reducing complexity by making a space, which is unintelligible to graphic reason, systematically investigable. Yet if the path through this space can be programmed and carried out, this hardly means that the intelligibility of the results of this path would necessarily be acquired. Indeed, calculation enables the automation of formal operations in such large quantity that the results are often difficult to apprehend; and to ascribe these results back to the rationality under whose direction the program was written often proves impossible. One becomes aware of this when observing the situation the Web offers today, and that which will probably be offered by the semantic Web of tomorrow: the internaut will be confronted with results obtained by interactions of heterogeneous, partly incomplete, or faulty sources, by means of complex protocols and calculations. Not knowing the sources enlisted, and not controlling the calculations, the internaut is confronted with a mass of results and documents which are not a priori intelligible to them. Calculation therefore presents a paradoxical situation, which consists of calling up a deficit of intelligibility, of understanding, in relation to a group of results and objects issuing from calculative rationality. It seems to us that this paradox emphasizes that our cognitive functions do not only confront us with a sensible nature offering us sensations; we are also confronted with a discursive nature, offering us not sensations, but inscriptions. These inscriptions, produced by our technology, and starting from intentions of communication, constitute no less a given, a discursive manifold, to take up an expression of Kantian inspiration, which we must make intelligible for ourselves. How can we make them intelligible? By leading them back to rational concepts and structures, which we already know and command. It is a question, therefore, of striving for a precise articulation between calculation and memory. Rhetoric as the Regulating Principle of the Numerical Discursive The problem consists of overcoming the discursive manifold by carrying out the pertinent “discursive” syntheses. In the absence of such syntheses the intellect is confronted by multiple representations, with-
Bruno Bachimont
378
out order or connection, leading it to a sense of disorientation. The discursive syntheses therefore have the task of unifying and joining the diverse given inscriptions, so that they make sense, so that they become intelligible. The intelligibility of the representations arises, thanks to a guiding thread, which can be placed between them. This guiding thread can be more or less rudimentary, going from the simple enumeration to narrative or argumentative articulation of the inscriptions. We propose considering four elementary procedures giving the discursive manifold a synthetic order. x
x
x x
Enumeration: the numerical or alphabetical order enables the assimilation of the aggregate of representations, and the transfer from one to the other, in an ordered and unconfused way. This order is rarely sufficient in itself, and a supplementary unity is often required: these are inscriptions with the same origin (the same author, issuing from the same work, the same stylistic form, the same finality – e.g. the psalms, which in the Middle Ages were learned from numerical grids, etc.). Narration: narration is the simplest manner of retaining heterogeneous elements for the purpose of joining them. As a procedure used abundantly in the management of knowledge (storytelling), narration enables the introduction, in an intelligible and graspable way, of a manifold. Argumentation: as an even more structured form, argumentation binds the manifold using reasoning, not a story. Perception: the discursive manifold is represented within the space of a support of a given inscription (e.g., the page that we represent to ourselves when we recall its content: a note, shopping list, etc.).
These procedures are not cognitive operations, in the sense of falling back on the natural function of the intellect. They correspond to methods elaborated by cultural traditions, which enable the individuals in a community to find themselves in the shared content, and the communal methods, of an investigation. These procedures are of a rhetorical origin: they correspond to strategies of the inventory which enable the traversing and recalling of a manifold, and the appropriation of the content, to invent the sense which it has for us. To make an inventory as a strategy of invention: this is the tradition that the rhetoric of antiquity and the Middle Ages15 has bequeathed us. A tradition which, among the 15
Cf. Mary Carruthers. The Book of Memory. A Study of Memory in Medieval Cul-
Formal Signs and Numerical Computation
379
five main parts which compose it (invention: finding the arguments of the discourse; elocution: finding the formulas of the discourse; disposition: articulating the elements of the discourse; action: restoring the discourse; memory: memorizing the discourse), articulates, in the narrow sense, invention and memory. I recall the discourse in order better to invent what I must do; I innovate because I repeat; I invent because I have a fund that I have assimilated. The discursive synthesis would therefore depend on procedures of a rhetorical nature, whose objective consists of mobilizing an old and memorized content, in order to elaborate a new discourse. Therefore these procedures would enable the traversal of the content presented by numerical systems in order to combine them into a new discourse, the understanding which I have of it. The modern internaut and the contemporary user of numerical technologies therefore find themselves in a situation analogous to that of someone who must study and assimilate a literary heritage: works whose origin or nature he sometimes hardly knows, which he then unifies in a network of common places, by means of which he constitutes a collective memory shared with others, and within which he finds the necessary material to express his thoughts and intentions. The conception of numerical systems must consequently take account of the intelligibility of its results, and think how, by rhetorical procedures of discursive synthesis, a reader or user can consult these results, as he would consult his memory or library. Using the paths forged by a tradition made of common-places, classical arguments, he can discover and appropriate the new and the unanticipated. Finally, computational reason must abolish calculation to make way for rhetoric: death to the algorithm, long live rhetoric, to plagiarize and contradict Victor Hugo. Not that one must abandon the possibilities of calculation and algorithm, but it is necessary to note the limits of their intelligibility, to mobilize rational not computational strategies, that is, argumentative, narrative, in short, rhetorical strategies. Physical Nature and Symbolic Nature In conclusion, we would like to come back to the remark made above: numerical systems remind us that nature is not only sensible, it is also symbolic. In other words, while the philosophical tradition often conture. Cambridge: Cambridge University Press, 1992 and idem. The Craft of Thought. Meditation, Rhetoric, and Making of Images. Cambridge: Cambridge University Press, 2000.
Bruno Bachimont
380 Sensible Nature
Symbolic Nature
Sensation
Signification
Spatio-temporally diverse
Discursively diverse
Confusion of senses
Conceptual disorientation
Objectivizing synthesis
Discursive synthesis
Schemas of comprehension
Rhetorical inventory
Physical engineering
Engineering of recognition
Fig. 2: Comparison of sensed and symbolic nature.
siders the concept as only the framing of the sensible by a form given by thought, the concept only being intelligible (sensé) when it enables the thinking of the sensible,16 signification only having the task of showing the giving of the sensible, the numerical tools show us that the symbolic inscriptions have their own autonomy. The inscriptions bring about inscriptions, independently of the sensible, whose conceptual formulation they could be. This casts a new light on the difficult and profound discussions concerning signification, knowledge, and being. By conferring on inscriptions their own dynamic, calculation reveals a symbolic nature. As the notion of an “experience of thought” suggests, thought as a manipulation of concepts constitutes an exteriority which experiences and investigates itself, which imposes and opposes itself, in short, which is capable of saying no to us. The numerical systems realize materially this exteriority of discursive thought, which we often investigate by means of simulation. Simulation is an authentic mode of experimentation on our inscriptions. However, as these inscriptions often treat the formalizing of sensible phenomena, they simulate the latter. But calculation deals with the efficacy of the concepts and not only with a simulation of phenomena (fig. 2). How can sensible and symbolic nature come together? Their coherence is not necessary, and an inscription is not deprived of meaning if it does not refer to a possible sensible experience. It is a symbolic experience. It remains for us to discover how sense, the sensible, and the intelligible (sensé) communicate and articulate themselves in our sensible and symbolic experience. The numerical systems help us to pose the 16
Of which Jocelyn Benoist reminds us in L’a priori conceptuel. Paris: Vrin, 1999.
Formal Signs and Numerical Computation
381
problem better, by adding the dispersion of the inscriptions to the confusion of the senses. 5. Conclusion It is easy to recognize in Hilbert’s program the optimism of the symbolic that Leibniz spoke of: algebraic symbolism makes a reasoning governed by certainty and assurance, and the avoidance of contradictions and paradoxes possible.17 This optimism, fruitful inasmuch as it enables the engendering of logical and mathematical formalisms, as well as their formal and automatist manipulation, nevertheless reaches impassable limits. If the mathematics (les mathematiques) are not reducible to a blind formalism themselves,18 then the problem, which has interested us here, is the use of formalisms in contexts where the signification of the utilized symbols is not determined in a rigorous way. In this case, one combines a technical possibility with a new power of invention, where the formal combinatorics engenders a productivity of thought. The intellect is exposed to symbolic configurations, engendered by a formal combinatorics, blind to sense and content, and alluding to thoughts which could never have been thought before. The algebra of symbols does not correspond to an algebra of thoughts, since the correspondence between symbols and thoughts is only approximate, and relies on a semiotic and linguistic interpretation. The interpretation of the resulting symbolic configurations is not predictable and cannot be reduced to a systematic research (to enumerate all possible cases of a formal system, to carry out all combinations of a given situation) where the systematic examination of symbolic configurations corresponds to the systematic investigation of a given situation. For this it lacks the rigorous correspondence between symbols and situations. The interpretation corresponds, rather, to a general heuristics. But this general heuristics equally gives rise to a dispersion of the symbolic inscriptions, in a profusion and confusion menacing their intelligibility. It is therefore necessary to question the validity of the utilized inscriptions, and their pretension to sense: a detour of a Kantian type allows us to discuss the foundation of sense attributed to them, as well as practising a critique of it. The proposed critique does not aim to disallow the use of the possibilities of formal techniques, but to suggest 17 18
Cf. Leibniz. Schriften. Vol. 7, 200. Hilbert. “Über das Unendliche.” Elliott Mendelson. Introduction to Mathematical Logic. Princeton: Van Nostrand, 1964.
382
Bruno Bachimont
a reasonable use of them, by relying, on the one hand, on a regulating principle based on rhetorical principles, and on the other, on a determining principle based on calculation. Their union opens an innovating and impassioned space for research, for which one begins to renew a variation of Leibniz’s optimism. Translation: Benjamin Carter WORKS CITED Bachimont, Bruno. Herméneutique matérielle et artéfacture. Des machines qui pensent aux machines qui donnent à penser. Ph.D. École Polytechnique, Paris, 1996. Bachimont, Bruno. “L’intelligence artificielle comme écriture dynamique. De la raison graphique à la raison computationnelle.” Au nom du sens. Ed. Paolo Fabbri and Jean Petitot. Paris: Grasset, 1999. 290-319. Belaval, Yvon. Leibniz critique de Descartes. Paris: Gallimard, 1960. Benoist, Jocelyn. L’a priori conceptuel. Paris: Vrin, 1999. Brague, Rémi. Wisdom of the World. The Human Experience of the Universe in Western Thought. Trans. Teresa L. Fagan. Chicago: University of Chicago Press, 2003. Carruthers, Mary. The Book of Memory. A Study of Memory in Medieval Culture. Cambridge: Cambridge University Press, 1992. Carruthers, Mary. The Craft of Thought. Meditation, Rhetoric, and Making of Images. Cambridge: Cambridge University Press, 2000. Goody, Jack. The Domestication of the Savage Mind. Cambridge: Cambridge University Press, 1977. Goody, Jack. The Logic of Writing and the Organization of Society. Cambridge: Cambridge University Press, 1986. Haugeland, John, ed. Mind Design. Cambridge: MIT Press, 1981. Hilbert, David. “Über das Unendliche.” Mathematische Annalen 95 (1926): 161-90. Kant, Immanuel. “Untersuchung über die Deutlichkeit der Grundsätze der natürlichen Theologie und der Moral.” Moses Mendelssohn. Abhandlung über die Evidenz in metaphysischen Wissenschaften, welche den von d. Kgl. Acad. d. Wiss. in Berlin auf d. Jahr 1763 ausgesetzten Preis erhalten hat. Berlin, 1764.. § 2, 73. Leibniz, Gottfried Wilhelm. “Projet de Préface à la science générale.” Die philosophischen Schriften von Gottfried Wilhelm Leibniz. Vol. 7. Ed. Carl Immanuel Gerhardt. Hildesheim: Olms, 1890. Leibniz, Gottfried Wilhelm. Opuscules et fragments inédits de Leibniz. Extraits des manuscrits de la Bibliothèque Royale de Hanovre. Ed. Louis Couturat. Paris: F. Alcan. 1903. Leroi-Gourhan, André. Gesture and Speech. Trans. Anna Bostock. Cambridge: MIT Press, 1993. Mendelson, Elliott. Introduction to Mathematical Logic. Princeton: Van Nostrand, 1964. Piérobon, Franck. Kant et les mathématiques. Paris: Vrin, 2003. Stiegler, Bernard. Technics and Time, 1: The Fault of Epimethus. Trans. Richard Beardsworth and George Collins. Stanford: Stanford University Press, 1998. Turing, Alan M. “On Computable Numbers, with an Application to the ‘Entscheidungsproblem.’” Proceedings of the London Mathematical Society 2.42 (1937): 230-65 and 2.43 (1937): 544-46.
DON IHDE
Art Precedes Science: or Did the Camera Obscura Invent Modern Science?
In November 2001, David Hockney’s Secret Knowledge. Rediscovering the Lost Techniques of the Old Masters1 was published with great fanfare. It made the claim that many artists from the Renaissance on used a now antique technology, the camera obscura, to make their paintings. Hockney, always somewhat controversial, had hit an issue. The release was followed within a few weeks by an overflow crowd conference at New York University with historians and critics responding to this thesis – claimed to be radical by Hockney. The dominant response was – well – “Heideggerian.” If Hockney is correct, then many revered artists could now be thought to have ‘cheated’ by copying an image produced by a device, a sort of drawing-by-the-lines approach taken as ‘inauthentic’ by the art crowd who implicitly favored hand and brush. Here is Heidegger: Human beings “act” through the hand; for the hand is, like the word, a distinguishing characteristic of humans. Only a being, such as the human, that “has” the word (mythos, logos) can and must “have hands.” . . . The animal has no hands, nor are hands derived from paws, claws, or talons . . . The hand has only emerged from and with the word . . . [For Heidegger, this becomes the privileged place for hand in handwriting] because the word, as the essential region of the hand, is the essential ground of being human . . . The word . . . symbolically inscribed . . . presented to vision is the written word, . . . script. As script, however, the word is handwriting. [But, handwriting, in typical Heideggerian dystopianism, gives way to technology, the typewriter] . . . It is not by chance that modern man writes “with” the typewriter . . . This “history” of the kinds of writing is at the same time one of the major reasons for the increasing destruction of the word. The word no longer passes through the hand as it writes and acts authentically but through the mechanized pressure of the hand . . . Mecha1
David Hockney. Secret Knowledge. Rediscovering the Lost Techniques of the Old Masters. New York: Putnam, 2001.
384
Don Ihde
nized writing deprives the hand of dignity in the realm of the written word and degrades the word to a mere means for the traffic of communication.2
Substitute the brush for the pen and you have much of the response to the Hockney thesis by many of the critics. When this publicity hype happened, I and a number of friends, equally engaged in the histories of science and technology and instrumentation, had some mutual chuckles over e-mail, another doubtlessly ‘inauthentic’ mode of communication. Catherine Wilson, the eminent philosopher and historian of microscopes, had already forewarned me of the Hockney thesis prior to the book’s appearance, after which we had a good laugh at the audacity of his proclaiming what most of the rest of us knew as common knowledge, i.e., that the Renaissance in both art and science was embodied through technologies, with the camera obscura being one favorite optical toy.3 My prized 1929 Encyclopedia Britannica – the same one which contained Husserl’s famous entry on phenomenology and Einstein’s on relativity – had an extensive article on the uses of the camera obscura as early as Alberti’s use in 1437. “The first practical step towards the development of the camera obscura seems to have been made by . . . Leon Battista Alberti, in 1437 . . . [referenced in] a biography, Rerum Italicarum Scriptores . . . It is stated that he produced wonderfully painted pictures . . . [via] a small closed box through a very small aperture.”4 And most historians knew of Da Vinci’s use of the camera to trace a cross from its image in 1450. And, what does one do about other technologies of drawing such as shown by Dürer and others (fig. 1)? No, Hockney did not rediscover the secrets of the Renaissance, he simply republicized them. What may have been forgotten by some art critics and historians is how fully technologized the Renaissance and 2 3
4
Heidegger quoted in Michael Heim. Electric Language. A Philosophical Study of Word Processing. New Haven: Yale University Press, 1987. 194-95. Cf. Catherine Wilson. The Invisible World. Early Modern Philosophy and the Invention of the Microscope. Princeton: Princeton University Press, 1995. In addition to being a history of the microscope, Wilson goes into detail about how it was taken by early modern philosophers, including Descartes and Locke. Regarding the familiarity of historically informed people concerning the camera obscura and its use by artists, in addition to a full article to this effect in the 1929 Encyclopedia Britannica, Peter Pollack’s The Picture History of Photography. New York: Henry N. Abrams, 1977 has a brief but detailed and illustrated history of art and the camera – and of its development into the photographic version of the camera. “Camera Obscura.” Encyclopedia Britannica. New York: Encyclopedia Britannica, 1929. Vol. 4, 658.
Art Precedes Science
385
Fig. 1: Albrecht Dürer. Artist drawing a Lute. Holzstich, 1525.
Early Modernity was. Might Galileo without his telescope be analogous to Caravaggio without his camera? First, then, what is the camera obscura? In simple terms, it is a very large ‘pinhole camera’; its literal meaning is dark room. The optic effect which the camera obscura captures was known in antiquity – probably known to Euclid and the Hellenic Greeks, but first thoroughly described in the revival and innovative expansion of this knowledge by the Arabic natural philosopher, Al Hazen, in 1037. Al Hazen’s Optics was the first major systematic treatise on light and optics and therein may be found a description of a camera obscura, which he probably used to observe a solar eclipse and which may have been the first anticipation of an early modern scientific use of the camera (fig. 2). The earliest European references to optics did not occur until 1270, with essays by Erazmus Ciolek Witelo and Roger Bacon – but by the Renaissance, the camera obscura was clearly one of the common optical toys of the time. Here is how it works: A light source, ‘outside,’ is directly cast through a small hole and then is produced as an ‘image’ on some flat surface ‘inside’ the darkened room. Carefully now note the effects and the ways in which the camera transforms an ‘object’ into an ‘image.’ The light source, here the sun (fig. 3), is produced in inverted form, and in two dimensions on the blank wall. (With the sun, these ef-
386
Don Ihde
Fig. 2: Description of a Camera obscura from Athanasius Kircher. Ars Magna Lucis et Umbrae (Rome, 1646).
fects might not be noticed, but if the image cast is one of an indirectly lit up object, the inversion and the two-dimensionality is easily perceived.) Our “Alberti” now can reproduce in a drawing this ‘automaticallyreduced-to-Renaissance-perspective’ image, emerge from the dark room and invert the drawing, and get the appreciative gasp of ‘verisimilitude’ his audience noted. One can see from this simple example how thoroughly so-called Renaissance perspective is ‘technologically’ produced. Its mathematization in one sense follows the optical effect. Now having noted the role the camera and related technologies played in the development and refinement of practices such as Renaissance perspective, the new way of seeing which occurs via ‘technological mediations’ illustrates the impact of optical technologies upon artistic practice. From this, we can equally note another indirect role played by the same technology. Al Hazen already noted in the eleventh century that the eye and the camera have certain analogous functions and shapes. Da Vinci later repeated this analogous observation: When the images of illuminated bodies pass through the small hole into a dark room . . . you will see on the paper all those bodies in their natural shapes and colors, but they will appear upside down and smaller . . . the same happens inside the pupil [italics mine].5 5
Da Vinci quoted by F. Lee Bailey. “Skull’s Darkroom. The Camera Obscura and
Art Precedes Science
387
Fig. 3: Illustrations emerge automatically from perspective presentation.
Here the camera begins to perform a metaphoric role. I say this because while there are some obvious analogical features: light through pupil, cast – as we now know much later than the Renaissance – upon the retina, etc., there are many others which are disanalogous. If the pupil is a ‘lens,’ it is a dynamic, variable one compared to the static one of the camera; the retina is a concave sphere in shape contrasting with the flat shape of the camera, etc., but metaphors emphasize similarities rather than dissimilarities. To this point, the earlier role of the camera in its relation to art practices has been noted. Already familiar and in use in the sixteenth century, even greater familiarity continued into subsequent centuries. And, if first used in art practice, the camera was also adopted into science practice. Indeed, Al Hazen had already used it as an instrument for observing solar eclipses and in the seventeenth century the ever inventive Galileo used a modification of the camera conjoined to a telescope as a helioscope as the means of observing sun spots – one of his major discoveries (at least for Europe – Chinese knowledge of sun spot activity had already been recorded several centuries earlier). Here, however, rather than review the various optical devices and experiments associated with seventeenth century science, I want to take the lifeworld familiarity of such technologies in a more philosophic and epistemological direction. Subjectivity.” Philosophy of Technology. Ed. Paul T. Durbin. Dordrecht: Kluwer, 1989. 66-67.
388
Don Ihde
One really large step in making the camera into a metaphoric device, into what I call an epistemology engine, occurs in the mid-seventeenth century. Both Descartes, in the Dioptics, and Locke, in the Essay on Human Understanding, specifically turn the camera into the very model for the production of knowledge, an engine of knowledge. Locke’s description is the most isomorphic: External and internal sensation are the only passages I can find of knowledge to the understanding. They are . . . the windows by which light is let into this dark room; for methinks the understanding is not much unlike a closet shut from light, with only some little opening left, to let in external visible resemblances, or ideas of things without . . . [these] resemble the understanding of man, in reference to all objects of sight and the ideas of them. [His tabula rasa was, of course, the blank wall of the camera.]6
Indeed, if we now use the camera as the model for knowledge production itself, one can get the outlines of both its terminology and its problems. Point by point, the sun is “external” reality, media res: the image cast, the idea or ‘thinking’ in the mind (fig. 4). The ‘subject,’ here invented as parallel to the ‘object,’ is inside the box of the body and sees only its own images/thoughts which correlate and represent that which is outside, or as it later becomes ‘external reality.’ But – and here comes problem one – if one’s ‘subjective knowledge’ is to be ‘true,’ it must correspond, like for like or isomorphically, with what is external. Thus Descartes plugs in his famous argument for God and his non-deceptive role guaranteeing the correspondence. One can also see here, at a glance, that the implied subject-as-homunculus (Einstein in the camera) also arises. What Descartes has done, is taken the metaphor of eye-camera, and expanded it into what I call an epistemology engine by making the metaphor stretch to: I=eye=camera. But Descartes (and Locke) employ a ‘cheat code.’ That is, the full description of what occurs inside the box is paralleled by a description of what occurs outside the box, thus implicitly presuming a point of view – were Descartes to be right and not be himself limited to a description as if he were truly only inside the box – which he could not have were knowledge limited to being inside the box. In short, Descartes occupies a position that allows him simultaneously to see both inside the box and outside the box, the position Donna Haraway calls the “god trick.” Descartes doesn’t really need God since he has already opted the god position! It was thus a technological model that shaped 6
John Locke quoted in ibid. 68.
Art Precedes Science
389
Fig. 4: Model of camera obscura according to René Descartes.
the basic outlines of early modern epistemology. We humans modeled our own processes upon one of our own inventions, a technological artifact. I could, and would like to go on, but the point I am making here is not one which engages in present debates about early modern epistemology – still vestigially virile in at least some disciplines today, although with the earlier Cartesian location of the implied ‘homunculus’ inside the pineal gland, is now re-located inside the ‘brain’ – it is instead intended to point to the way in which a technology plays an important role, one which even makes it into the very model of knowledge production itself. To show this I want to take the camera obscura beyond both its Renaissance and seventeenth century settings and relate its later modifications to the same knowledge production activities. The camera obscura was a tool which through an amazing number of proliferations of technological development and modification can be shown to be one of the most versatile springs of scientific instrumentation, but it, by becoming a metaphor, also became the epistemology engine which justified and explained how scientific knowledge itself worked. And while this may in some ways fall into the Heideggerian observation that technology is ontologically prior to science, it is an observation which is much more concrete, empirical, and historical than Heidegger could be. Nor can I here yield to the very strong temptation to show how early modern epistemology no longer works well, particularly in science, and
390
Don Ihde
with that follow moves to more contemporary equivalents of replacement epistemology engines – which I suspect all of you would recognize in the roles of computer and internet processes which are rapidly being metaphorized into cosmology itself – but instead I want to show how, in a much more concrete and phenomenological sense, the camera obscura invented science in different and recognizably actual ways: (1) The optical effect captured by the camera goes back to antiquity, probably known to Euclid, but first fully described by Al Hazen, who also seems to have used it to watch an eclipse – just as we do today with cardboard versions. Galileo also adapted a camera device, a helioscope attached to his telescope, which cast an image showing the sun to have spots – his report of this was the first publication to get him into trouble with the Vatican. (2) But, its first major social use was in art, as indicated, and the camera was a sort of ‘automatic’ process by which ‘Renaissance perspective’ was produced. The camera ‘automatically’ reduced three-dimensional objects to two-dimensional projections, with ‘perfect’ proportionality. Note that this is a transformation of a ‘natural object’ into what we often call an ‘image.’ If one extends this insight, one can see that the way of seeing associated with Renaissance perspective, mathematized as ‘geometrical projections,’ invents a ‘geometrical method’ which precedes by a couple of centuries what is usually taken as the birth of early modern science. I also want to note that this early use of the camera is limited to its isomorphic imaging possibilities, i.e., to make ‘realistic’ depictions. If the style of vision that is associated with ‘Renaissance perspective’ is one of ‘realism,’ it owes much of this to camera production, and in this Hockney is justified. (3) This same isomorphic possibility – as I am sure you have already anticipated – becomes, in the nineteenth century, another ‘camera,’ this time a photographic one. There is no example in the history of technology that I know to have exceeded the speed of dissemination than that occupied by photography. The process, actually invented by Joseph Niépce in 1826, was bought out by Louis-Jacques-Mandé Daguerre and publicized in his best seller on how to build a camera in 1839. Scientists began using this process almost immediately, as in this first shot of the gibbous moon in 1840. Only two years later (1842) the astronomer, J.W. Draper, photographed a solar spectrum. And while one can easily see that the photographic camera is a reincarnation, a simple modification of the earlier obscura, one should not pass by without noting its different technological transformations of objects into images. Early photography was still, it ‘fixed’ its images, and for science this was impor-
Art Precedes Science
391
tant because the fixed image could be returned to again and again for deeper and deeper analysis. Note, too, in early form, how the technology non-neutrally produced instrumentally selected objects. Exposure time was such that at first only still objects – landscapes and architecture mostly, could be photographically imaged. That also partly accounts for the ‘stiff poses’ of early portraiture. (I cannot follow here the proliferation of other possibilities that photography took up with faster shutter speeds and the like, but my general point is that photographic camera technology, like all technologies, is non-neutrally selective and transforms all its objects into what we now usually call images.) (4) A third reincarnation of the camera obscura might be missed, because what counts for imaging in this version takes a very different shape, a non-isomorphic shape. I refer to spectroscopy, which got its start via Isaac Newton. Newton, like Galileo, an actual optics maker, wrote in 1666, “I procured me a triangular glass prism to try therewith the celebrated phenomenon of colors . . . Having darkened my chamber and made a small hole in my windowshuts to let in a convenient quantity of the sun’s light, I placed my prism at its entrance, that it might be thereby refracted to the opposite wall.”7 In short, another camera obscura, simply now replacing the hole – which very often was also lensed – with a prism. The result was that the image cast was that of the rainbow spectrum. Newton himself did not go on to invent the spectroscope – but from these experiments he did derive a theory of color and brilliantly found a way to overcome the color aberrations associated with any refractive telescope over 30 power – Newton invented the reflecting telescope which re-focuses the different frequencies of colored light into a single location by way of a parabolic mirror. It was not until more than a century later that chemists, still playing with the prism version of the camera, found that by replacing the round hole of the classical camera with a slit, they could get a far better image, one which was found to have dark lines rather than fused colors; these are called Fraunhofer lines after a chemist by that name produced them in 1814. These were, again later, finally discovered to be unique chemical signatures such that one could ‘read’ the chemical signatures of the sun and stars. Today, spectroscopy has proliferated into thousands of types of instruments, which in a parallel with optical telescopy, now can exceed the ranges of light itself. Spectroscopy allows us to ‘read’ nature’s 7
Isaac Newton, letters, quoted in “Isaac Newton and Robert Hooke, Dispute on the Nature of Light.” Galileo’s Commandment. Ed. Edmund Blair Bolles. New York: W.H. Freeman, 1997. 184f.
392
Don Ihde
bar codes. But the instruments are recognizable variations upon the camera obscura. (5) Nor does the story end there. My colleague, Bob Crease writes for Physics World and he recently polled physicists about what they took as the ten most beautiful experiments in the history of physics, and the Science Times, September 24, did an article about these. Out of the ten, three are about light and all are variations upon camera obscuralike instruments. Newton’s prism version is listed and ranks number 4. A little over a hundred years later, Thomas Young (1803) makes another modification on the camera, instead of a lensed hole or a diffraction slot, he uses a double slit. He cut a hole in a window shutter, covered it with a thick piece of paper punctuated with a tiny pinhole and used a mirror to divert the thin beam that came shining through. Then he took a ‘slip of a card’ . . . and held it edgewise in the path of the beam, dividing it in two . . . The demonstration was often repeated over the years using a card with two holes to divide the beam . . . [these are now called] double-slit experiments.8
And, voilà, yet another camera variant. Physicists ranked Young’s experiment number 5. Rank number 1 stretches the double slit a bit, but only because instead of light, a beam of electrons is passed through a double-slit device yielding the famous quantum ‘wavicle’ or particle/ wave phenomenon so dear to contemporary physics – the actual experiment occurred in 1961 by Claus Jonnsen at Tubingen. The camera is brought into the late twentieth century – and I sometimes wonder what a triple slit might do? At the very least, these variations, or technological trajectories from the early, basic camera obscura, allow the sciences to see, ‘read,’ and probe into the microscopic and macroscopic phenomena, which expressed, are the knowledges produced by science and its menagerie of camera obscurae. These are the productions that ‘invent’ much modern science. I want, however, to leave you not simply with this epistemic history as it were, but to suggest something more provocative: Each variant upon the camera, isomorphic, non-isomorphic, and today often constructive, produces visual displays, images, which are what are seen and read by the trained scientific hermeneut. But the scientific object is, precisely in its strongest sense, that transformed phenomenon presented in and through the instrumental technology. Nature does not come to modern 8
Quoted by George Johnson in “Here They Are, Science’s 10 Most Beautiful Experiments.” New York Times (Tuesday, September 24, 2002).
Art Precedes Science
393
science raw or naked, but technologically transformed. The natural object must be transformed into the ‘scientific object’ to be observed. That is the deeper secret of the camera obscura. Science, however ‘realist,’ or however ‘socially constructed,’ is also technologically constructed.
WORKS CITED Bailey, F. Lee. “Skull’s Darkroom. The Camera Obscura and Subjectivity.” Philosophy of Technology. Ed. Paul T. Durbin. Dordrecht: Kluwer, 1989. Bolles, Edmund Blair. “Isaac Newton and Robert Hooke, Dispute on the Nature of Light.” Galileo’s Commandment. Ed. idem. New York: W.H. Freeman, 1997. Encyclopedia Britannica. Vol. 4. New York: Encyclopedia Britannica, 1929. Heim, Michael. Electric Language. A Philosophical Study of Word Processing. New Haven: Yale University Press, 1987. Hockney, David. Secret Knowledge. Rediscovering the Lost Techniques of the Old Masters. New York: Putnam, 2001. Johnson, George. “Here They Are, Science’s 10 Most Beautiful Experiments.” New York Times (Tuesday, September 24, 2002). Pollack, Peter. The Picture History of Photography. New York: Henry N. Abrams, 1977. Wilson, Catherine. The Invisible World. Early Modern Philosophy and the Invention of the Microscope. Princeton: Princeton University Press, 1995.
THOMAS F. GIERYN
Instrumentalities of Place in Science and Art
I have two stories about schools, one a school of art, the other a school of science. The first story is about the New York School of Abstract Expressionists – a school of painters. The second is about the Chicago School of urban sociologists – a school of social scientists. The first story is set in the period from the 1930s to the mid-1950s, just after the peak of surrealism and just before pop art. The second is set in the period from the 1890s to the 1930s (and so it comes out of sequence). The first story has famous players in it: Jackson Pollock, Willem De Kooning, Robert Motherwell, Barnett Newman, Clyfford Still, Mark Rothko, Adolph Gottlieb, Philip Guston, Franz Kline, Arshile Gorky. The second story also has players, but maybe they are famous now only among sociologists and urban theorists: Albion W. Small, Robert E. Park, William I. Thomas, Ernest W. Burgess, Louis Wirth, William F. Ogburn. My two stories are neither diffuse nor complete but focused and partial – they speak only about the places where effervescent cultural production once happened, two American cities (for art, in New York; for science, in Chicago). After I tell these two stories, I shall draw out some sociological themes about the emplacement of art and science – about the geographical and architectural situatedness of collective action that yields legitimate understandings (that is: art accepted as worthy; science accepted as true). Instruments used for scientific and artistic work have a place. They exist at some particular spot in the universe – bounded, spatially delimited. They are located amid assemblages of buildings, streets, tools, supplies, cafés and maybe trees. Such assemblages, such places, themselves become instruments for the production of science and art – consequential in ways not fully captured by the separate tools of production in the hands of an artist or scientist. Places gather up at some location not just things but also people, organizations and social institutions – and all of that gets brocaded in layers of narration and interpretation to create a cultural meaning associated with “being there.”
Instrumentalities of Place in Science and Art
395
How did New York matter for the Abstract Expressionists? How did Chicago matter for the urban ecologists? To what extent is the beginning – and the end – of these two schools formed by the cities for which they are named? Could New York have been the sine qua non for its Abstract Expressionist painters, as Chicago might have been for its urban sociologists? Two analytic questions are asked. First, why is the production of art and science geographically concentrated in recognizable places (and not dispersed more evenly throughout the countryside)? Second, what are the consequences of this clumping or clustering for the value of scientific and artistic œuvres? How does the place of provenance of a scientific discovery or artistic work matter for its worth (that is, how it is received and understood by diverse audiences and consumers)? The analysis suggests a cultural boundary distinguishing art from science: in art, place-based labels for circles of artists and their works help to establish value and secure reputations; in science, the worth of claims is achieved by their detachment from immediate geographical and material circumstances of production. In art, a place of provenance is celebrated; in science, it is transcended and erased – and these different processes are epitomized when artists and scientists are labeled according to the city or region or territory where they work. My New York story: Abstract Expressionism might have taken root somewhere else, or (less probably) nowhere in particular – scattered widely through geographic space, like the air. It is appropriate that the term “Abstract Expressionism” – used to describe the activities of a stylistically disparate but geographically centered group of painters – appeared for the first time in The New Yorker magazine, in March 1936, in an article by Robert Coates. Now, Abstract Expression and New York are inseparable, no matter where the artists may live or where their paintings might hang. Pollock, De Kooning, Motherwell, and Rothko1 comprise the “New York School” – a label found in all the art history books to signal where these artistic activities happened, and a label that has an instrumentality in its own. But why at this place in particular? Why did Abstract Expressionism happen in New York? The artists themselves tell us what the place felt like to them, in the 1930s and 1940s. New York, New York . . . exhila1
On the routine expungement of Black, homosexual, and female participants from the New York School, cf. Ann Eden Gibson. Abstract Expressionism. Other Politics. New Haven: Yale University Press, 1997.
396
Thomas F. Gieryn
rating, romantic, energizing, motion, action, spontaneity, freedom, anxiety, complexity, expansive, expressiveness (emotions not unlike the paintings they made).2 Chaotic and neurotic, where an artist can (and must) become himself. “Unknown, uncozy, and not small scale.”3 Its huge population enables both immersion and seclusion, the anonymity of getting lost in a crowd or retreating to a garret. New York has no tradition other than a relentless annihilation of traditions (as W.H. Auden said).4 Paris is velvet, silk, soft clinging woolen stuff. New York is made of cutting metal . . . Paris beats the measure; New York moves in rhythms. Paris has its gait; New York walks. Paris has secrets; its ear is open to obscure lives that are bound up in it – that will not be heard about until later; in New York everybody is famous, everybody figures on the first page.5
New York already had everything and so there was nothing left to do – a challenge for these artists, then, to create something for which there was no anticipation or want. The extremes are ever present and on display: glittering wealth on Fifth Avenue, bums on The Bowery. New York is not-Europe, but neither is it the rest of America – but instead always regarded with suspicion in the provinces and heartlands as too foreign, too cosmopolitan, too unconventional. And, maybe, too demanding in its standards of quality: the City does not suffer fools, and “If you can make it there, you can make it anywhere.” Young artists, like Clyfford Still from North Dakota, came to New York “for a time to be part of a movement, to share experiences with other painters, and to win acknowledgment in the great metropolis.”6 New York was full of the institutional infrastructures from which a budding bunch of painters could appreciate the best art that came before them – learn from it, mock it or respect it, and then do otherwise. New York, the confluence. Its universities gathered scholars such as art historian Meyer Schapiro at Columbia – Motherwell left Harvard to come to New York to study with Schapiro.7 Its museums gathered and 2
3 4 5 6 7
Cf. Dore Ashton. The New York School. A Cultural Reckoning. New York: Penguin, 1972; Bernard Rosenberg and Norris Fliegel. The Vanguard Artist. Chicago: Quadrangle Books, 1965. Street impressions by de Kooning, cf. Ashton. New York School. 4. Quoted in Ashton. New York School. 136. Michel Seuphor. “Paris New York 1950.” Modern Artists in America. First Series. Ed. Robert Motherwell et al. New York: Wittenborn Schultz, 1951. 118-22. Ashton. New York School. 34. Cf. Nancy Jachec. The Philosophy and Politics of Abstract Expressionism, 19401960. Cambridge: Cambridge University Press, 2000.
Instrumentalities of Place in Science and Art
397
showed both old and new art from Europe: Impressionists to Surrealists (although the Museum of Modern Art had little interest in works by non-European artists until deep into the forties). The Gallantine Collection was at New York University, downtown, from 1927 to 1942, “with its large and well-selected group of cubist paintings by Picasso, and scores of other abstract works from many countries.”8 New York’s patrons, like Peggy Guggenheim, gave temporary refuge and gallery space to wartime exilic artists such as Mondrian, Leger, Ernst, Tanguy, Dali, and Albers – in New York but not of New York (Parisian vestiges). Guggenheim’s Art of This Century, a gallery at 30 West 57th Street,9 provided wall space for the segue from Europeans to Americans – Motherwell, Baziotes, Rothko, and others had showings there in the mid-1940s. New York, the magnet – a magnet whose immense mass drew so much creativity from so far.10 Fodder for the painters: previous generations of art and artists close at hand, creating a tolerable ambivalence (admire it, exceed it), but demanding excellence. “Take New York out of this model of the movement’s origination, . . . and there would have been no movement – it fused there.”11 New York, the “nerve center.”12 New York . . . this is synecdoche, of course.13 Abstract Expressionists were not scattered throughout the five boroughs, but tightly packed in lower Manhattan, downtown, centered on the Village, Eighth Street, 8 9 10
11
12 13
Ashton. New York School. 30f. Ibid. 121; Martica Sawin. Surrealism in Exile and the Beginning of the New York School. Cambridge: MIT Press, 1995. vi. And not just painters, but musicians as well, such as John Cage, Morton Feldman, Earle Brown, and David Tudor. Cf. Steven Johnson. “Introduction. A Junction at Eighth Street.” The New York Schools of Music and Visual Art. Ed. idem. New York: Routledge, 2002. 1-15. David Thistlewood. “Historicity and Mythology in American Abstract Expressionism.” American Abstract Expressionism. Ed. idem. Liverpool: Liverpool University Press, 1993. 1-39. Some scholars have challenged the “Jerusalemic significance of New York,” preferring to treat the place as mere happenstance whose significance has been concocted by scholars engaged in myth-making. Cf. Serge Guilbaut. How New York Stole the Idea of Modern Art. Chicago: University of Chicago Press, 1983. William C. Seitz. Abstract Expressionist Painting in America. Cambridge: Harvard University Press, 1983. 4. “New York” also stands in for the rest of the United States, as one of its capitals. The rise of Abstract Expressionism is emplaced in the nation as well as in the city, and a sizable bibliography has grown up (for example) around the implications of Pollock & Co. for American cold war politics: cf. Guilbaut. How New York Stole; Jachec. Philosophy and Politics of Abstract Expressionism.
398
Thomas F. Gieryn
and Tenth Street. Greenwich Village was fertile ground for growing the next avant-garde, the next artistic vanguard. Its reputation as the capital of bohemia was well-established from the 1920s, by writers such as Malcolm Cowley: demi-monde, jazz, poetry, homosexuality, café society (a little like Paris).14 A small town, “little trees and crooked streets” that did not quite conform to Manhattan’s grid (like Montmartre), notso-tall buildings, even an outdoor art exhibition (where Franz Kline first appeared) – but a small town always “all edge,” challenging and resisting bourgeois mores, an “alien outpost” below 14th Street.15 The Village tolerated these artists – but as painters, they were not instantly welcomed into the creative counterculture then dominated by writers. Was painting art? Was it serious? Could it wrestle with timeless insecurities? The Village nurtured Abstract Expressionists even as it forced them to intellectualize what they were doing: the downtown radical literati made words as important as oil on canvas, and so the Abstract Expressionists talked and talked and talked, endlessly into the night. It is also no geographic accident that the place where Abstract Expressionism took root was not far from Union Square –“where comrades meet” (an advertisement for the New China Cafeteria on Broadway near 14th Street, a “convenient, low-priced eating place which the artists frequented, comrades or not, and indulged in what had become a New York obsession: talk.”16) Union Square was the site of demonstrations and protests during the depressing days of the 1930s – socialists, communist and labor activists of all kinds gathered there (and so did some of the artists, comfortable for a time with the leftist politics they found there). An Artists Union was formed – like a cell of red proletariat, but engaged in cultural production – and the magazine Art Front was its mouthpiece. Some eventually famous Abstract Expressionists were supported by the Works Progress Administration, and therein enjoyed the “sense of having found each other.”17 More ambivalence, more tension: Union Square politics asked for engagement with the harsh realities of unemployment and poverty, art in the service of ideology, not unlike what some of the social realists were doing. Diego Rivera, David Alfaro Siqueiros, Thomas Hart Benton passed through New York, with their murals about injustice and exploitation. Some 14 15
16 17
Cf. Ashton. New York School. 20ff. Harold Rosenberg. “Tenth Street. A Geography of Modern Art.” Discovering the Present. Chicago: University of Chicago Press, 1973. 100-09. “Alien outpost” is from Ashton. New York School. 49. Ashton. New York School. 53. Ibid. 44.
Instrumentalities of Place in Science and Art
399
Abstract Expressionists like Gorky (who had a studio at Union Square) flirted with radical leftist politics in the 1930s – who could resist, with misery all around them, there, below 14th Street, in their faces and noses? – but love did not last. The New York School, especially as the Depression gave way to War and then peace in the fifties, sought deeper realities than the political and economic – subconsciousness itself, transcendent expressions of meanings that floated away from the Marxist moorings they eventually left behind at Union Square. “Gorky would be shouted down at Artists Union meetings when he tried to discuss aesthetic questions.”18 Art for art’s sake became their slogan, not “workers of the world, unite!” Art critic Harold Rosenberg wrote that “a de Kooning canvas is as unrestricted as Union Square,” and (much later) de Kooning himself would reflect on his roots: “I come from 36 Union Square.”19 Lower Manhattan offered more than ideas – what once were factories and warehouses became studios, for living in and painting in, relatively inexpensive, and available for the taking. Loft rats, they were called, inhabiting dingy “cold-water” walk-up flats that offered abundant open spaces and sunlight filtered through a century of grime accumulated on windows and skylights.20 Tanguy the surrealist lived among them, on Waverly Place in the Village. De Kooning was on 12th Street and Fourth Avenue, enjoying the sacred squalor of a space apart from (and above) all the literature and politics and wretchedness swirling around outside. Studio space was sometimes shared: Bradley Walker Tomlin with Motherwell at Astor Place.21 Chance meetings on the street, impossible unless they were all at New York together, led to innovation and spontaneity: “One evening in the winter of 1940-1941, Baziotes brought Jackson Pollock over to Kamrowski’s studio, and the three artists began experimenting with quick drying lacquer paint that Baziotes had bought at Arthur Brown’s art supply store.”22 Charles Duits, a Harvard undergraduate at the time, describes the scene: 18 19 20
21 22
An observation from Peter Busa, quoted in Sawin. Surrealism. 91. Harold Rosenberg. Willem de Kooning. New York: Harry N. Abrams, 1973. 13. De Kooning is quoted in Sawin. Surrealism. 413. Clement Greenberg wrote in a celebrated 1939 essay “Avant-Garde and Kitsch:” “The fate of American art is being decided by young people, few of them over forty, who live in cold-water flats and exist from hand to mouth.” Quoted in Guilbaut. How New York Stole. 160. April Kingsley. The Turning Point. The Abstract Expressionists and the Transformation of American Art. New York: Simon & Schuster, 1992. 167. Cf. Sawin. Surrealism. 168, 334.
400
Thomas F. Gieryn
In New York one could always encounter either the members of the Surrealist group or other well known artists such as Chagall, Mondrian, and Leger, or those whose reputations were still to be made, Jackson Pollock, Gorki [sic], David Hare, Baziotes, Motherwell.23
Another artist looked back on those days: “Here in New York you just step out on the street, and you can either take it or leave it, go visit people or stay home. The freedom’s fantastic.”24 The black and white humongous calligraphy of Franz Kline mirrors the loft, in its contrast between sooty black exteriors and whitewashed walls inside. Was he painting “New York”? Or was he just painting in New York, with New York, of New York? Abstract art is not exactly representational, so the Village did not provide scenery – rather, the subject matter of those paintings was the painter’s subjectivity. “Gottlieb’s images are insignia of remoteness, of a continent or cosmos of the mind as distant as possible from the sign systems or twentieth-century New York.” “No one could mistake it [the Village] for an aesthetic creation . . . [it is a] neutral zone.”25 However private and personal the lofts and studios might have been,26 they were never far away from other lofts and studios – where different artists were making different paintings that were not so different. The cult of the solitary artistic genius is belied by the communalism of these same artists, who competed among themselves but depended upon each other too for survival, for approval, for beacons, for reassurance, for business. “All the artists know each other and see each other. This does not mean that there is no competition, but it is not ferocious.”27 It was a school, after all, not independent autonomous selfemployed operators; and they needed a brand name, a signature style, to differentiate themselves from other art-commodities. The artists needed each other to be close at hand: “The experimentation among younger American artists was being carried on largely without benefit of a support system or the possibility of feedback except from other artists.”28 Tension again: both the individualistic and communal im23 24 25
26
27 28
Ibid. 223. Rosenberg and Fliegel. Vanguard Artist. 18. Harold Rosenberg. “Gottlieb.” Adolph Gottlieb. Paintings 1950-1971. London: Marlborough Gallery, 1971. Not paginated; Harold Rosenberg. “Tenth Street.” 103 and 105. “The ‘cold-water flats’ they occupied became famous sites of creative solitude.” Caroline A. Jones. Machine in the Studio. Constructing the Postwar American Artist. Chicago: University of Chicago Press, 1996. 7. Seuphor. “Paris New York 1950.” 122. Sawin. Surrealism. 333.
Instrumentalities of Place in Science and Art
401
pulses are inscribed geographically and architecturally, in places where the right people could meet and talk, and then retreat and concoct, and then reconvene to judge and praise, and then withdraw to worry or exhale. One artist said: “You have the drench yourself in the current scene, then get away.”29 And some could never bear to leave.30 Attachment to this place became an identity. The American kandy-kolored kool-aid writer Tom Wolfe wrote The Painted Word in 1975, and Abstract Expressionism comes off as a sophisticated shell game in which (by the mid-fifties) audiences, patrons, critics, gallery operators, and museums are hoodwinked into believing that this art is good (worth buying).31 What made it good? Did New York make it good art? Tom Wolfe describes the Abstract Expressionists as inhabiting a “cénacle” (Latin: Coenaculum; English: the Upper Room) – and the allusion works perfectly. Cénacle is the place where Jesus gathered with his disciples to perform transubstantiation in the Eucharist, where the Last Supper might have happened, and where the Spirit of Christ returned on Pentecost to empower his followers to spread the Good News. It is, for Christians everywhere, a kind of “truth-spot.”32 Outside theology, cénacle has come to mean a club – both the chosen people gathered and the places where they do so. There were many cénacles for the Abstract Expressionists, enabled by the architectural surrounds made available by earlier generations of New Yorkers. Around 1940, European emigrés and upstart American painters gathered at the unusually capacious loft of experimental filmmaker Francis Lee, “in an industrial building on Tenth Street . . . we lived in cramped, cold apartments and had no cafés to meet in so we went to Francis’s loft several times a week.”33 Hans Hofmann’s art school on Eighth Street, which lasted from 1938 to 1958, where the Expressionists assembled to learn the European traditions so that they could unlearn them. Atelier 17 at the New School for Social Research, transplanted from Paris by printmaker Stanley William Hayter, where “thanks to the cooperation of the Maltese maintenance staff it ran until the small hours of the morning . . . and artists [Motherwell, Baziotes, Rothko, Gottlieb, Pollock] began making a habit of dropping in to work
29 30 31 32 33
The unnamed artist is quoted in Rosenberg and Fliegel. Vanguard Artist. 16. Katharine Kuh. The Artist’s Voice. New York: Harper and Row, 1960. 153. Tom Wolfe. The Painted Word. New York: Bantam, 1975. 44ff. Thomas F. Gieryn. “Three Truth-Spots.” Journal of the History of the Behavioral Sciences 38 (2002): 113-32. The artist Roberto Matta Echaurren is quoted in Sawin. Surrealism. 110f.
402
Thomas F. Gieryn
or talk.”34 The Waldorf Cafeteria (open all night) on Sixth Avenue near Eighth Street, and later the Cedar Tavern on University Place and 8th, along with Romany Marie’s, the San Remo, Stewart’s and Riker’s, bars and coffee houses: surrogates for Parisian cafés,35 talk, talk, talk. “Tenth Street has its public lounges that serve as gossip exchanges and gardens of romance; its cocktail- and evening-party circuits with their snobberies; its mutual aid.”36 The Club, an “agora”37 for art and artists only38 and predictably on Eighth Street, where (a little later, during the fifties) lectures about art were heard by increasing throngs of now-famous painters and their wannabees, represented a definite attempt to bind together a community . . . Many a youthful provincial heard about the Club and about the Cedar, and came to New York to have a chance to drink with the big boys, to call out ‘Hi, Jack,’ or ‘Hi, Bill,’ or ‘Hi, Franz,’ and to feel a part of something not yet defined that was happening in the arts.39
In 1951, the Ninth Street show, the “first salon of the New York School [which] signified . . . the existence of a community of artists, like-minded chiefly in their dedication to an open-ended approach to painting,”40 where patrons and customers could buy art with an added authenticity – because it was made right in the neighborhood, like tourist art that one buys not in the airport but in the distant village. Leo Castelli helped hang the art for the Ninth Street show, anticipating his own downtown gallery featuring the Abstract Expressionists – and although this show was little noticed in the media, the momentum was there. With critics Harold Rosenberg and Clement Greenberg to decipher its meanings, and an increasing number of entrepreneurial dealers (Sam Kootz, Sidney Janis, Charles Egan, Betty Parsons) willing to ratify and invest, the School had arrived where it had never left: Greenwich Village, NYC. Cénacles are places for believers to find each other and assemble, to witness miracles, to learn and perpetuate the canon, to define and pro34 35 36 37 38
39 40
Ibid. 152. Kingsley. Turning Point. 72. Cf. Sawin. Surrealism. 139. Rosenberg. “Tenth Street.” 107. Dore Ashton. “Introduction.” William C. Seitz. Abstract Expressionist Painting in America. Cambridge: Harvard University Press, 1983. xvii. “In the original charter for the Club, for instance, the members stipulated that the meetings should exclude the world of commerce and politics: no dealers, no critics, no Communists.” Evidently women were initially excluded too (a barrier quickly overcome by Ruth Abrams and Elaine de Kooning), a reminder that cénacles exclude as well as include. Gibson. Abstract Expressionism. xxv. Ashton. New York School. 198-99. Cf. Kingsley. Turning Point. 72f. Sawin. Surrealism. 413.
Instrumentalities of Place in Science and Art
403
ject what holds them together – as a religious sect, as an artistic school. The cénacles below 14th Street defined the boundaries of Abstract Expressionism, as they defined the boundaries of “good” modern art. Hofmann’s school, the Cedar Tavern, the Club were spatializations of membership, separating those who were inside Abstract Expressionism from those outside. Studio 35 (on Eighth Street, of course), where an art school called “The Subjects of the Artist” began in 1948, was the place where the artists gathered in 1950 to define themselves: “It had no platform; it had no poet-spokesman; and no one knew what to call it, let alone how to describe it.”41 In the cénacles, one could see and hear about what others were painting, each in his own unique way – but enough alike (in principle and purpose) that they could be stamped with the legitimacy of the “New York School,” and purchased thus with confidence. You had to be there.42 The network was in place now: “the milieu in which these two groups of artists connected with each other – that is, art dealers, curators, patrons, critics, and a diversified intellectual community.”43 But, then, only spatialized ironies. The Downtown that gave Abstract Expressionists their locale and identity gets invaded by Uptown: monied gallery-owners who go slumming for good art, art with the imprimatur of Eighth Street – as the artists themselves put up a weak resistance to this invasion, eventually submit, and become rich and famous (and “New York School” becomes like Mercedes, no longer a place were “it” was happening, but a safe investment).44 And the Cedar Tavern is invaded by younger hungrier artists-in-chrysalis – hordes of them (Jasper Johns, Robert Rauschenberg, Andy Warhol), who seek to define their own legitimacy and worth by inhabiting the same places where Pollack sat and talked, or Rothko. Like a synagogue that becomes the flashpoint for anti-Semitism, the cénacles of Lower Manhattan themselves become targets of artists seeking their own idiom different from the one developed by the old artists who now only haunt 41
42 43 44
Ibid. 379f. My description of The Club also draws on Robert Goodnough’s account in Robert Motherwell et al., eds. Modern Artists in America. 9. Perhaps this undefinable quality of Abstract Expressionism led Motherwell to say in 1951 (curiously, I think): “The recent ‘School of New York’ – a term not geographical but denoting a direction – is an aspect of the culture of modern painting.” Quoted in Guilbaut. How New York Stole. 204. Claude Cernuschi. ‘Not an Illustration, but the Equivalent.’ A Cognitive Approach to Abstract Expressionism. Cranbury: Associated University Presses, 1997. 14. Sawin. Surrealism. 196. For sales figures, cf. Guilbaut. How New York Stole. 91.
404
Thomas F. Gieryn
these places. The originals have left for the Island – the Hamptons and Montauk. The school is dead … long live the school! In the end, Cedar Tavern fails to exclude impostors, ingrates, heretics, connivers – and, by the time it meets the wrecking ball, Abstract Expressionism is “New York” only in name, and Pop Art is on the ascent. The New York scene has become “too full of duties and manners.”45 “The artists’ Tenth Street will not deteriorate; it will be extinguished. It will be swallowed in the yawn of a steam shovel. Its future is – an excavation.”46 My Chicago Story. Curiously, perhaps, there are few schools of thought in the social sciences named after the cities where they grew up (the Frankfurt School of critical theory comes readily to mind47) – and a remarkably high number of such instances involve Chicago, which has well-known and city-labeled schools of political science and economics and, of course, the Chicago School of Sociology. In fact, there are even two “Chicago Schools of Sociology,”48 though I limit my attention to the first: a group of social scientists gathered on the Midway at the University of Chicago between 1892 and the early 1930s, whose research centered on the structures and processes of urban life – “urban ecologists.” The Chicago School49 is called that because of a tight coupling of the city and the University that bears its name. The University of Chicago is in and of the city of Chicago in ways that surpass the local embeddedness of any other major American research university. Part of the connection is that city and University grew up together – and both enjoyed the benefits of “novelty effects.” Chicago was still a relatively small prairie town at the end of the American Civil War, but with rail45 46 47 48
49
Cf. Rosenberg and Fliegel. Vanguard Artist. 19. Rosenberg. “Tenth Street.” 109. Martin Jay. The Dialectical Imagination. A History of the Frankfurt School and the Institute of Social Research, 1923-1950. Boston: Little, Brown & Co., 1973. The “second” Chicago School of Sociology takes off in the late 1950s, and continues on through the 1970s, built on the foundations of symbolic interactionist theories and qualitative methods provided by Everett Hughes and Herbert Blumer, and including central figures such as Erving Goffman, Howard Becker, Anselm Strauss, Eliot Freidson, and Joseph Gusfield. Cf. Gary Alan Fine, ed. A Second Chicago School? Chicago: University of Chicago Press, 1995. My emphasis on the Chicago urban sociologists means that little is said of the “first generation” symbolic interactionists and social psychologists such as George Herbert Mead. Exactly what makes it a “school” is under debate. Cf. Andrew Abbott. Department and Discipline. Chicago Sociology at One Hundred. Chicago: University of Chicago Press, 1999. 21ff.
Instrumentalities of Place in Science and Art
405
roads, “kaleidoscopic”50 immigration, industrial booms, and a spectacular lakefront location roughly in the middle of America’s “manifest destiny” march across the continent, Chicago became a huge worldclass metropolis by the beginning of the twentieth century.51 Even the Great Chicago Fire of 1871 was less a calamity than a mere bump in the road – an opportunity to build anew all over again, phoenix-like. As Carl Sandburg put it in his poem “Chicago:” “Laughing the stormy, husky, brawling laughter of youth.”52 “Chicago [is] . . . comparatively young,” and “in a mere sixty years, rose from nothing.”53 The University of Chicago was also new, having opened its doors only in 1892 – a babe among the long-established elite universities of the East Coast, with whom the U of C would soon compete for intellectual supremacy. And differences between Chicago and Harvard or Yale can be traced in part to their age: the U of C could develop with fewer tugs from the past, and far fewer pretensions. Among other things, from the very start, the University saw itself as implicated in the city of Chicago – it was not just an accidental location, but a spot chosen to improve them both. The Department of Sociology at the University of Chicago was especially close to the city around the campus. That city offered a propitious context for the various research agendas that soon emerged: “Materially, the city is the superior social object.”54 Max Weber visited in 1904, on his way to the World’s Fair in St. Louis, and compared Chicago to “a man whose skin had been peeled off and whose intestines were seen at work.”55 Because of its relatively recent growth in population, Chicago displayed the fundamental ecological processes of early twentieth century American society: the shift from farm or village to city, increasing racial and ethnic diversity (as Blacks moved north and Europeans crossed the Atlantic in huge numbers), a lumpy melting pot 50
51 52 53
54
55
Homer Hoyt. One Hundred Years of Land Values in Chicago. Chicago: University of Chicago Press, 1933. XI. On Chicago School studies of race and ethnicity, cf. Stow Persons. Ethnic Studies at Chicago, 1905-45. Urbana: University of Illinois Press, 1987. William Cronon. Nature’s Metropolis. Chicago and the Great West. New York: Norton, 1991. Quoted in Martin Bulmer. The Chicago School of Sociology. Chicago: University of Chicago Press, 1984. XIV. Frederic M. Thrasher. The Gang. A Study of 1313 Gangs in Chicago. Chicago: University of Chicago Press, 1927. 147; Marco D’Eramo. The Pig and the Skyscraper. Chicago, A History of Our Future. London: Verso, 2002. 256. T.V. Smith and Leonard D. White, eds. Chicago. An Experiment in Social Science Research. Chicago: University of Chicago Press, 1929. 245. “Jerusalemic significance” of Chicago. Cf. D’Eramo. Pig and the Skyscraper. 8. Bulmer. Chicago School. 23.
406
Thomas F. Gieryn
where some new arrivals assimilated and thrived while others have yet to do so, rapid industrialization resulting in both “gold coasts” (unprecedented wealth) and “slums” (unprecedented poverty). For Park, the new big city of Chicago “not only releases aspects of human nature which would otherwise remain suppressed, remolding both the character and essence of the human being, but also brings forth completely new ‘varieties.’”56 Chicago (the city) was an ideal subject matter for these urban ecologists. Along with social change came social troubles – and these also contributed to the tight connection between Chicago School sociology and its home-place. The problems of economic inequity were in the face of the urban ecologists, and the following description of the South Side today is not far from what it was like eight decades ago: “the area of Hyde Park, where the University of Chicago is located, is a tiny enclave of privilege literally besieged by some of the city’s most derelict Black neighborhoods.”57 From its beginning, the University of Chicago – seat of can-do pragmatist philosophy (John Dewey) – viewed learning as justified not just by greater knowledge but progressive social reform. Chicago sociologists took to their city with a zealous commitment to make it better – through careful objective research that could help to stem the tide of crime, injustice, misery, and social disorganization of all sorts. The city was not only the place to observe the most consequential social processes of the day, but also the place to manage them. How much this reform could be pursued alongside an increasingly “scientific” sociology (objective, value neutral, truth for its own sake) was a fundamental tension in the Chicago School.58 It is no accident that Jane Addams, founder of Hull House and key figure in the use of “settlement houses” to improve the lives of poor immigrant populations, was never welcomed as a player in the U of C Sociology Department – and she is remembered more often as a “social worker” instead.59 56 57 58
59
Rolf Lindner. The Reportage of Urban Culture. Robert Park and the Chicago School. Cambridge: Cambridge University Press, 1996. 60f. D’Eramo. Pig and the Skyscraper. 397. Henrika Kuklick. “Chicago Sociology and Urban Planning Policy.” Theory and Society 9 (1980): 821-45; Steven J. Diner. “Department and Discipline. The Department of Sociology at the University of Chicago, 1892-1920.” Minerva 13 (1975): 514-53; Abbott. Department and Discipline. 31. Lester R. Kurtz suggests that “all of the works [of the Chicago School] were within the corporate-liberal technocratic tradition.” Evaluating Chicago Sociology. Chicago: University of Chicago Press, 1984. 27. Mary Jo Deegan. Jane Addams and the Men of the Chicago School, 1892-1918. New Brunswick: Transaction Books, 1988.
Instrumentalities of Place in Science and Art
407
The dense connections between University and city were organizational as well. Members of the Sociology Department joined the City Club, Union League Club, and the Chicago Urban League (Park was its first president) – civic organizations of businessmen, politicians, and professionals (with power). Fruits of the sociologists’ work insinuated their way into the bureaucracy of Chicago’s local government. Throughout the 1920s, Vivien Palmer assembled a “Social Base Map” for Chicago, an early anticipation of today’s computerized GIS (Geographical Information Systems). The city was divided into 75 natural areas, then broken down into 300 neighborhoods, and social data were collected for each (crime rates, property values, land use, racial/ethnic mix, and so forth).60 Evidently, some of the boundaries created in this immense data-gathering project are still used today by city officials tracking, for example, the delivery of municipal services like garbage pick-up or schooling. In 1924, sociologist Ernest W. Burgess was chosen to head up the Chicago Census Committee, charged with the enumeration (and analysis) of everybody who lived within the city limits. This tradition of local social involvement was the soil that nurtured the urban studies of the Chicago school of sociology and made the relationship between the university and city different from that in New York or Boston.61
The Chicago School also grew because of the robust infrastructure provided by the University of Chicago itself. John D. Rockefeller, Sr. gave $35 million over two decades to get the ball rolling, and that amount rolled much farther a century ago than it would today. Later, the Carnegie Corporation and the Laura Spelman Rockefeller Memorial provided copious funds for University initiatives in the social sciences – specifically, the Local Community Research Committee, an organizational shell through which most of the famous Chicago case studies were carried out, like The Gold Coast and the Slum, The Jack-Roller, The Ghetto, The Gang, The Taxi-Dance Hall.62 The word “Local” in the name of this Committee is not incidental: the Rockefeller Memorial fund spent its money to get sociologists out into Chicago, to study the city of Big 60 61 62
Bulmer. Chicago School. 119. Ibid. 25. Harvey W. Zorbaugh. The Gold Coast and the Slum. Chicago: University of Chicago Press, 1929; Clifford R. Shaw. The Jack-Roller. A Delinquent Boy’s Own Story. Chicago: University of Chicago Press, 1930; Louis Wirth. The Ghetto. Chicago: University of Chicago Press, 1928; Thrasher. The Gang; Paul G. Cressey. The Taxi-Dance Hall. Chicago: University of Chicago Press, 1932.
408
Thomas F. Gieryn
Shoulders, and to glean from their research some policies for reform and progress (benevolent capitalism).63 The Department of Sociology itself provided an institutional infrastructure well-suited to the nurturance of a budding “school.” Here, one could find a “dense, highly integrated, local network of teachers and graduate students carrying out a program of research in one city centered around common problems.” In addition, there was “an infrastructure and simple research organization to support it . . . coupled with outside financial support for the work which was being done.”64 The Department was led by able and possibly charismatic men: Albion Small at the start, then William I. Thomas, later Park and Burgess, and near the end, Wirth and Ogburn (all of them “big names” in the discipline, then and now). The Department at Chicago was the first in the world to offer graduate training in the new field of sociology, produced more new Ph.D.’s than any other Department,65 established the discipline’s first scholarly periodical – The American Journal of Sociology66 (still the most prestigious), enjoyed the benefits of a new book series on sociology at the up-and-coming University of Chicago Press, and – in 1921 – Park and Burgess published their Introduction to the Science of Sociology, the first textbook in sociology and enormously influential.67 Chicago the Department dominated the scholarly means of production in sociology for thirty years, just as Chicago the city dominated the observations and models of its practitioners. And yet it was still a cozy place – only a handful of faculty, with most of the case studies carried out by graduate students, for their dissertations. A Gemeinschaft of social scientists amid the Gesellschaft of Chicago the city. I have not yet uncovered the equivalent of a Cedar Tavern for Chicago School sociologists, but I’m sure there was one. Or: maybe they worked all the time, with no occasion for drink or idle chatter. A graduate student from the 1920s said that everybody essentially lived in the Department’s offices, and they were cramped (initially, Park and Burgess shared a single office in the East Tower of Harper Library).68 The environment got much better in 1929, with the opening of the Social Sci63 64 65 66 67 68
Cf. Bulmer. Chicago School. 139; Kurtz. Evaluating Chicago Sociology. 62. Bulmer. Chicago School. 1, 7. Kurtz. Evaluating Chicago Sociology. Chapter 7. On the history of the American Journal of Sociology, cf. Abbott. Department and Discipline. Chapters 3, 4. On the textbook, cf. Robert E.L. Faris. Chicago Sociology 1920-1932. Chicago: University of Chicago Press, 1967. Chapter 3. Bulmer. Chicago School. 95.
Instrumentalities of Place in Science and Art
409
ence Research Building at 1126 East 59th Street (more Rockefeller money, and still the Department’s home today). The instrumentalities of place are vivid in the design of this building: it has no large classrooms for undergraduate teaching, but just seminar rooms; the offices were not clustered by discipline into departments, but rather interdisciplinarity was spatialized (sociologists, political scientists, economists, and anthropologists were strewn throughout the floors).69 No fewer than six offices were allocated for journal editorships; and – importantly – there were laboratories here, not just for archaeology and physical anthropology (bone rooms), but for sociology: a card sorting room, four statistical labs (where Ogburn ruled the roost), drafting rooms, map display rooms, archive storage facilities. There was also, in the Social Science Research Building, a common room where tea was served daily – maybe that was the Cedar Tavern. Edward Shils remembered that social scientists “saw their colleagues when they came in the morning, they saw them when they went to lunch.”70 With this full head of steam, it is a wonder that the Chicago School of Sociology could not survive the 1930s intact – and I think that the city of Chicago had as much to do with the decline as with its thirty years of success. Just as the Chicago sociologists struggled to limn the differences between pure research and applied, they also struggled to decide which methodologies were best suited for understanding the social. Indeed, the department’s orientation to sociology was “varied and eclectic, and its strength lay in this diversity.”71 The School may be best remembered today for its qualitative ethnographic field methods – participant observation, mainly, but also personal life histories and interviews. Vivien Palmer wrote a field manual to instruct new sociologists on how to conduct research on the streets of Chicago: immerse yourself, be patient, surrender, and gain “intimate contacts of life” . . . and walk, walk, walk.72 Make “spot maps” to show the spatial patterns in Chicago’s social life, like Snow’s cholera map – only for crime, delinquency and other vices. Robert Park took his students on walks through Chicago, not for exercise but for science.73 The intimacy between University and city extended to the methodologies of its sociologists: up-close and personal. Get to know the par69 70 71 72 73
Quoted in ibid. 193. Ibid. 197. Ibid. 3. Vivien M. Palmer. Field Studies in Sociology. A Student’s Manual. Chicago: University of Chicago Press, 1928. 41. Cf. Bulmer. Chicago School. 97.
410
Thomas F. Gieryn
ticulars of this place, what makes this neighborhood different from any other (the sights, the smells). Yield to the givenness of social life on its own terms – in the streets and tenements – without imposing upon them heavy conceptualizations. Park especially encouraged his students “to leave the study rooms and go out on to the uncertain terrain of ‘real life’” – “nosing around.”74 Sociology in the wild, in the raw – but so rich, because (after all) this place of observation is Chicago (where all the makings of modern society are splayed open for view). The city as field-site. Being there.75 For all that the city of Chicago gave to these urban sociologists, it was also their undoing – at a time (in the thirties, and after) when the discipline desired to make itself look scientific, more reliable and valid, more objective. Quantitative methods soon became the rage, and the goal shifted from detailed case studies of the particulars here or there to generalizations and universalizations true everywhere. Chicago (as object of study) became, now, in this new methodological regime, a liability – whereas before, it was a unique asset for those sociologists gathered on the Midway, on 59th Street. Rather than walking the streets of Chicago, the ambition became hypothesis-testing using variable-based models and drawing on systematic statistical surveys, as in Ogburn’s research on the 1928 presidential elections.76 The point was not to understand Chicago, or one of its neighborhoods, but to examine the effects of religious belief and alcohol-tolerance on voting patterns – statistical relationships that could be measured anywhere (and it did not matter where). Chicago became a source of data, not a uniquely equipped place to observe and engage the workings of society first-hand. “Nosing around” “does not sound very serious when measured by the standards of the scientific community.”77 All that research – on Chicago, in Chicago, of Chicago, about Chicago, for Chicago. But could the findings be generalized to anywhere, or (at least) to any big industrial American city? The Chicago School strained to make their perfect field-site (with all of its unique and local advantages) into a laboratory – standardized so that it could yield knowledge that was not place-specific, but universal instead. It was not enough that the Department built laboratories inside the Social Science 74 75
76 77
Lindner. Reportage of Urban Culture. 1. Cf. Bulmer. Chicago School. 45 and 97. The core of Chicago School sociology is “that one cannot understand social life without understanding the arrangements of particular social actors in particular social times and spaces.” Abbott. Department and Discipline. 196. Cf. Bulmer. Chicago School. 175. Lindner. Reportage of Urban Culture. 2.
Instrumentalities of Place in Science and Art
411
Research Building at 1126 East 59th Street. To escape the provincialism of their findings, the entire city of Chicago must itself become a laboratory. “The city . . . is in a very real sense a laboratory for the investigation of collective behavior.”78 City as lab – now, an analytic space rather than a place with sounds and smells (and a truth-spot of a different sort) – rendered with statistical correlations rather than observations from the street or from the person, statistical correlations that have no particular location. Park himself aimed “beyond concrete analysis at generalizations,” and Wirth sought “sociological laws.”79 The desire to universalize Chicago – even after the city had given its sociologists so much as a singularity – is evident in what has been called “the most famous diagram in sociology.”80 Ernest W. Burgess’s “concentric zone model” is a description of how different land usages are geographically arrayed.81 The model is ecological, designed to understand better the processes of invasion and succession, as new immigrant groups invade neighborhoods, compete among themselves, upset the old equilibrium, only later to give way to some new geographic accommodation (sometimes assimilation). One could see the dynamic city through this conceptual model – its structure and processes. But which city? In one of the diagrams offered by Burgess, the referent is plainly Chicago: there is the shore of Lake Michigan, here is the Loop (downtown), there is Little Sicily and the Underworld to the Northwest, the Ghetto and Deutschland to the West, the Bright Light Area and Chinatown to the South. Overlaid on top of what is recognizably Chicago are a nested set of concentric circles, labeled Zones I through V. From the core, they are: Loop, Zone in Transition, Zone of Workingmen’s Homes, Residential Zone and Commuters’ Zone. It is still Chicago, but now theorized into different spaces (zones) for different functions. Burgess offers a second diagram: the theory is now denuded of Chicago’s particulars, just an abstract set of doughnut zones, created by faceless placeless processes, but generalizable – applicable to any city anywhere. Burgess admits lamely: “Neither Chicago nor any other city 78
79 80 81
Robert E. Park and Ernest W. Burgess. The City. Chicago: University of Chicago Press, 1925. 22. The concept that Chicago must become a laboratory in order for it to be studied scientifically comes from Bruno Latour. “Circulating Reference. Sampling the Soil in the Amazon Forest.” Pandora’s Hope. Cambridge: Harvard University Press, 1999. Chapter 2. Lindner. Reportage of Urban Culture. 107; Wirth. The Ghetto. 6. Mike Davis. Ecology of Fear. Los Angeles and the Imagination of Disaster. New York: Random House, 1998. 364. Park and Burgess. The City. 51-55.
412
Thomas F. Gieryn
fits perfectly into this ideal scheme.”82 What has been lost? Chicago the field-site, with all those peculiarities and idiosyncrasies that insisted on a slow walk and patient familiarity.83 All of this is given up in the service of universal facts, and “Chicago” becomes a laboratory for social science en plein air, stripped of whatever might have made it a real place – just a space for testing theories and models. No longer was Chicago the uniquely right tool for the job. Like my story about the New York School of Abstract Expressionists, the Chicago School of sociology also learned that whatever the city gives, the city can take away. Place can be an instrument for – and sometimes against – the production of worthy art and science. These stories about the emplacement of two schools of cultural production evince similarities between art and science – but also important differences. Abstract Expressionists and urban ecologists are both geographically clumped: members of each school are clustered tightly together in New York or in Chicago. The two schools are labeled by the cities where they worked, and place-names are remembered and invoked as we try to make sense of those artists and scientists, and what they did. There is likely to be a reason for this pattern, found out when one looks into the social conditions that are salutary for transforming individual creativity into collective achievement. And yet, the place of provenance matters differently for art and science, in this sense: the value of paintings by Abstract Expressionists depends forever upon their rootedness in New York, while the value of findings, concepts and theories by urban ecologists depends instead upon their detachment from Chicago. Art and science work with distinctive cultural logics. The enduring worth of a painting is established by the particulars of its making – the artist, the period, the place of production, the style or genre. The generalizing logic of science requires that such particularities be denied in the pursuit of truths that – if valid – must transcend the local conditions of their making. Cultural production is circumstantial. Artistic and scientific practices depend quite literally on “things standing around” – close at hand, then and there.84 These creative activities clump geographically for the same reason that business activities clump: “agglomeration efficiencies,” as the 82 83 84
Ibid. 51f. Cf. Kurtz. Evaluating Chicago Sociology. 26. Cf. Karin D. Knorr-Cetina. The Manufacture of Knowledge. Oxford: Pergamon, 1981. 33; Andrew Pickering. The Mangle of Practice. Time, Agency, and Science. Chicago: University of Chicago Press, 1995. 59.
Instrumentalities of Place in Science and Art
413
economist Alfred Marshall called them.85 For example, global capitalism is not really that at all. Manufacturing and managerial activities are concentrated only here or there, and huge spaces on the globe are empty of them altogether. Global capitalism creates “global cities” like New York, London, and Tokyo – where multinational corporations station themselves so that they can find ready-at-hand the accoutrements of profit-making (such as: law firms who specialize in international tariffs; information technology experts who can fix the most complicated computer system; reliable phone and delivery services). Neither fast airplanes nor even faster digitally mediated communications can eliminate completely the costs of transit. Moving goods, people, ideas remains costly, and so (by implication) does the geographic dispersion of mutually dependent operations and auxiliaries. Ironically, it is even beneficial to have capitalist competitors nearby: to watch their every move, to steal their employees, to battle together against common external enemies (like government regulatory agencies). Having the wherewithal of business close at hand is efficient: steelmaking once clumped in Pittsburgh and the Ruhr; high finance now is in New York or maybe Frankfurt. So it goes with art and science: they clump, geographically. Both Abstract Expressionists and urban ecologists benefited from a legacy of place-specific meanings that were associated with New York or Chicago well before these artists and scientists arrived. The Bohemian atmosphere of Greenwich Village (or the radical ambience of Union Square) and the “newness” of Chicago (or the can-do optimism of its pragmatist philosophers) encouraged these two schools to challenge received wisdom, to experiment, to innovate. Moreover, their accomplishments depended upon institutional and material circumstances that were also not of their making – just “available” in the cities called home: world-class art museums and cheap cold water-flats contributed something to Abstract Expressionism, just as smelly stockyards, a perfumed Gold Coast, hordes of immigrants from everywhere and Rockefeller largesse contributed something to urban ecology. As places, New York and Chicago assembled close at hand what was needed for dayto-day cultural production: Arthur Brown’s art supply store was as necessary as the card-sorting laboratory in the Social Science Research Building. New York and Chicago were big enough to serve as magnets, drawing into their confluence people just passing through – emigré surrealists, Max Weber – who left their marks on the art, and on the sci85
For my understanding of “agglomeration economies,” I rely on Saskia Sassen. The Global City. 2nd ed. Princeton: Princeton University Press, 2001.
414
Thomas F. Gieryn
ence. New York and Chicago gathered up the people who made Abstract Expressionism and urban ecology possible. Downtown Manhattan was an “art world,” consisting of many more than just the heralded painters.86 Art is collective activity requiring diverse social roles for its accomplishment: the painter of course, but also the patron, the curator, the critic, the audience. New York assembled all those people, shoulder to shoulder – just as Chicago assembled a “science world.” The Chicago School was not just Park and Burgess, but also visionary administrators at the University of Chicago, dozens of graduate students who studied the city for their dissertations, technical assistants like Vivien Palmer, and even Jane Addams as a foil. It is possible to disaggregate all of these circumstantial conditions that were so vital for the two schools to succeed – and treat each factor (say, a Peggy Guggenheim or a Laura Spelman Rockefeller) as an autonomous independent cause of cultural effervescence. But why? Plainly, the whole is greater than the sum of its parts. A place – like New York or Chicago – agglomerates these disparate circumstances so that they combine, so that they merge spatially into a force for artistic or scientific production far more consequential than any one of them alone. Was New York the sine qua non for its Abstract Expressionist painters? Was Chicago an “obligatory passage point”87 for its urban ecologists? Absolutely yes. But was it imperative that Abstract Expressionism happened in New York, or urban ecology in Chicago? “Geographic determinism” has little merit (as if there were something essential about New York or Chicago, making it necessary that art or science flourish only under those unique conditions.) Both art and science have flowered in less hospitable places. An acclaimed school of American impressionist painters came together in a backwoods, virtually inaccessible and underpopulated hamlet in upland Indiana during the early twentieth century – the “Brown County” School. Thousands of scientists developed fifteen research institutes and pursued cutting-edge physics during the 1960s in Siberia (of all places), near Novosibirsk – Akademgorodok.88 However, neither Brown County nor Siberia did much on their own – as places – to bring together clumps of artists or scientists. T.C. Steele and other American impressionists got to know each other in the more salutary 86 87 88
Howard S. Becker. Art Worlds. Berkeley: University of California Press, 1982. Bruno Latour. Science in Action. How to Follow Scientists and Engineers through Society. Cambridge: Harvard University Press, 1987. 150. Lyn Letsinger-Miller. The Artists of Brown County. Bloomington: Indiana University Press, 1994; Paul R. Josephson. New Atlantis Revisited. Princeton: Princeton University Press, 1997.
Instrumentalities of Place in Science and Art
415
surrounds of fin-de-siècle Paris, and were later networked into the Chicago big city art market; in the same way, mathematician Mikhail Alekseevich Lavrentev could not have enticed any physicist or biologist to Novosibirsk without Khrushchev’s cold war Kremlin politics. The location of a school of cultural production may indeed be arbitrary – but only if something serves as a catalyst to bring together artists or scientists into co-presence. Often the catalyst is a city – the place itself draws them together, and this is surely so for the New York and Chicago schools. The reputations, institutions, organizations, philanthropic benefactors, and material surrounds drew artists to New York in the thirties and forties, and urban sociologists to Chicago in the teens and twenties. And gradually, more artists and scientists were attracted to these two places simply by each other – younger artists gathered at the Cedar Tavern to be around Motherwell, just as students came to the Windy City to sit at the feet of Robert Park. Once assembled for a time in Mecca (or Jerusalem), the tight network of true believers can then disperse themselves without sacrificing the place-based experiences that connected them. Abstract Expressionists fled to the Hamptons only after they had become friends and rivals amid the dense milieu of Union Square and the Village; students of Park and Burgess took the Chicago School to jobs at universities throughout the land. One could, I suppose, see these two schools as merely “networks” of linked individuals with no common ground of location – but what got them linked up in the first instance? Once a nascent school of art or science begins to coalesce at some particular location, heavy “path dependencies” exert themselves: temporally sequenced events gradually amass (in an iterative way) a foundation that tightens the bond between cultural production and a place. Pollock’s first gallery showing in New York made it more likely that the second exhibition of an Abstract Expressionist would also take place there, which increases the odds of the third, and so forth. The first dissertation published under the aegis of Park and Burgess increased the probabilities of a second Chicago School thesis, third, and many more. Place captures the momentum of a developing school, and materializes it in a juggernaut of museums, studios, lofts, galleries, laboratories, common rooms where art and science is talked about, practiced, scrutinized, criticized, and made public. The institutionalization (and routinization) of artistic and scientific practices is geographic and architectural, at least in part. Place does two more things to stimulate cultural production in a school. First, at the level of individual creativity, some places are good
416
Thomas F. Gieryn
to think with (or, at least, better than some other places for getting creative juices to flow). Neurophilosopher Andy Clark has introduced the concept of “cognitive scaffolding” to describe the material world outside the brain, onto which humans off-load creative thinking.89 That is, we design and build places in order to routinize some cognitive tasks – or maybe to stimulate or evoke non-routine thinking (creativity). I don’t know enough yet about the actual architecture of the lofts and studios of the Abstract Expressionists (the windows, the walls, the ceilings, the square footing) – or about the labs at 1126 East 59th Street in Hyde Park, but evidence from other studies of where creative people work suggests that Clark may be correct.90 Second, creative work seems to depend upon chance encounters with others – unexpected face-to-face meetings that can provide the moment for spontaneous conversation or perhaps an unplanned mutual endeavor. Co-presence at some location (residing in the Village, or Hyde Park) is a prerequisite for such unpredicted meetings – and cénacles or common rooms are built explicitly to encourage them.91 Still, the cultural logics of art and science diverge from each other, and place is also useful for understanding the cultural boundaries between them. Place of provenance adds value to works of art but subtracts it from works of science (no matter how vital those places might be for the production of those works). For Abstract Expressionists, hanging out on Eighth Street (or Tenth) defined membership in what would become the New York School. “Being there” ratified the artist and the œuvre as genuine and authentic (put differently, artists who pursued Abstract Expressionism without ever stepping foot in Greenwich Village during the 1940s were much less likely to have their paintings canonized). Artistic reputations are embedded in a place, presumably because living in Downtown Manhattan embued the artist with skills honed in the highly-charged collaborative but competitive scene, 89 90
91
Andy Clark. Being There. Putting Brain, Body and World Together Again. Cambridge: MIT Press, 1997. Jill Krementz. The Writer’s Desk. New York: Random House, 1996; David Seidner. Artists at Work. Inside the Studios of Today’s Most Celebrated Artists. New York: Rizzoli, 1999; Peter Galison and Emily Thompson, eds. The Architecture of Science. Cambridge: MIT Press, 1999; David N. Livingstone. Putting Science in Its Place. Chicago: University of Chicago Press, 2003. The relationship between building design and the propensity for chance encounters in science and engineering is explored by Thomas J. Allen. Managing the Flow of Technology. Cambridge: MIT Press, 1977. 236ff.; cf. Peter Galison on “trading zones” where scientists happen to meet. Image and Logic. A Material Culture of Microphysics. Chicago: University of Chicago Press, 1997. 803ff.
Instrumentalities of Place in Science and Art
417
enabling the artist to achieve what gets defined as “good art.” Patrons and purchasers look to place as a “brand name” contributing to the lasting value of the canvasses on their walls. Even as paintings by Rothko or de Kooning get hung in museums and galleries in Berlin, they remain “of New York” – interpreted as from there, in catalogues and art history texts, taking meaning and value from where they originated. Even as some masterpieces seek transcendent and universal value, their locational roots still matter: they speak to all humanity from some place. No art is from nowhere. It is different in science, where detachment from the place of production is a sign of objectivity, and a measure of the validity (or value) of reality-claims. Scientists pursue a “view from nowhere,” the “freefloating intelligentsia” is celebrated, and the cosmopolitan trumps the local. Laboratories become “placeless places,” designed to be hermetically sequestered from environs outside their walls, and – through their architectural uniformity – interchangeable and mobile.92 Science fears the parochial, the private, the particular, and the promiscuous – a set of sins absolved only by extracting facts and theories from the locales in which they were born, and reinstalling them on a universal plane (true everywhere). Embedding a theory of the city in the Social Science Research Building on 59th Street, or in Hyde Park, or in a particular Gold Coast or slum, or even in Chicago, stigmatizes it. Validity depends upon distanciation from the source, upon the transit of the theory to other places – to see if it fits not-Chicago, not-Hyde Park. Attachment of scientific claims to specific places compromises their value. We memorialize places where great art was made, venerate them, visit them. We submerge places where great science was made, beneath replications, confirmations, extensions, and applications.
92
Thomas Nagel. The View from Nowhere. Oxford: Oxford University Press, 1986; “free-floating intelligentsia” is from Karl Mannheim. Ideology and Utopia. New York: Harcourt Brace Jovanovich, 1936; Robert K. Merton. “Patterns of Influence. Local and Cosmopolitan Influentials.” Social Theory and Social Structure. New York: Free Press, 1968. Chapter 12; “placeless places” is from Robert E. Kohler. Landscapes and Labscapes. Chicago: University of Chicago Press, 2002.
418
Thomas F. Gieryn
WORKS CITED Abbott, Andrew. Department and Discipline. Chicago Sociology at One Hundred. Chicago: University of Chicago Press, 1999. Allen, Thomas J. Managing the Flow of Technology. Cambridge: MIT Press, 1977. Ashton, Dore. The New York School. A Cultural Reckoning. New York: Penguin, 1972. Ashton, Dore. “Introduction.” William C. Seitz. Abstract Expressionist Painting in America. Cambridge: Harvard University Press, 1983. XV-XXVI. Becker, Howard S. Art Worlds. Berkeley: University of California Press, 1982. Bulmer, Martin. The Chicago School of Sociology. Chicago: University of Chicago Press, 1984. Cernuschi, Claude. ‘Not an Illustration, but the Equivalent.’ A Cognitive Approach to Abstract Expressionism. Cranbury: Associated University Presses, 1997. Clark, Andy. Being There. Putting Brain, Body and World Together Again. Cambridge: MIT Press, 1997. Cressey, Paul G. The Taxi-Dance Hall. Chicago: University of Chicago Press, 1932. Cronon, William. Nature’s Metropolis. Chicago and the Great West. New York: Norton, 1991. Davis, Mike. Ecology of Fear. Los Angeles and the Imagination of Disaster. New York: Random House, 1998. Deegan, Mary Jo. Jane Addams and the Men of the Chicago School, 1892-1918. New Brunswick: Transaction Books, 1988. D’Eramo, Marco. The Pig and the Skyscraper. Chicago, A History of Our Future. London: Verso, 2002. Diner, Steven J. “Department and Discipline. The Department of Sociology at the University of Chicago, 1892-1920.” Minerva 13 (1975): 514-53. Faris, Robert E.L. Chicago Sociology 1920-1932. Chicago: University of Chicago Press, 1967. Fine, Gary Alan, ed. A Second Chicago School? Chicago: University of Chicago Press, 1995. Galison, Peter. Image and Logic. A Material Culture of Microphysics. Chicago: University of Chicago Press, 1997. Galison, Peter and Emily Thompson, eds. The Architecture of Science. Cambridge: MIT Press, 1999. Gibson, Ann Eden. Abstract Expressionism. Other Politics. New Haven: Yale University Press, 1997. Gieryn, Thomas F. “Three Truth-Spots.” Journal of the History of the Behavioral Sciences 38 (2002): 113-32. Guilbaut, Serge. How New York Stole the Idea of Modern Art. Chicago: University of Chicago Press, 1983. Hoyt, Homer. One Hundred Years of Land Values in Chicago. Chicago: University of Chicago Press, 1933. Jachec, Nancy. The Philosophy and Politics of Abstract Expressionism, 1940-1960. Cambridge: Cambridge University Press, 2000. Jay, Martin. The Dialectical Imagination. A History of the Frankfurt School and the Institute of Social Research, 1923-1950. Boston: Little, Brown & Co., 1973. Johnson, Steven. “Introduction. A Junction at Eighth Street.” The New York Schools of Music and Visual Art. Ed. idem. New York: Routledge, 2002. 1-15. Jones, Caroline A. Machine in the Studio. Constructing the Postwar American Artist. Chicago: University of Chicago Press, 1996. Josephson, Paul R. New Atlantis Revisited. Princeton: Princeton University Press, 1997.
Instrumentalities of Place in Science and Art
419
Kingsley, April. The Turning Point. The Abstract Expressionists and the Transformation of American Art. New York: Simon & Schuster, 1992. Knorr-Cetina, Karin D. The Manufacture of Knowledge. Oxford: Pergamon, 1981. Kohler, Robert E. Landscapes and Labscapes. Chicago: University of Chicago Press, 2002. Krementz, Jill. The Writer’s Desk. New York: Random House, 1996. Kuh, Katharine. The Artist’s Voice. New York: Harper and Row, 1960. Kuklick, Henrika. “Chicago Sociology and Urban Planning Policy.” Theory and Society 9 (1980): 821-45. Kurtz, Lester R. Evaluating Chicago Sociology. Chicago: University of Chicago Press, 1984. Latour, Bruno. Science in Action How to Follow Scientists and Engineers through Society. Cambridge: Harvard University Press, 1987. Latour, Bruno. Pandora’s Hope. Cambridge: Harvard University Press, 1999. Letsinger-Miller, Lyn. The Artists of Brown County. Bloomington: Indiana University Press, 1994. Lindner, Rolf. The Reportage of Urban Culture. Robert Park and the Chicago School. Cambridge: Cambridge University Press, 1996. Livingstone, David N. Putting Science in Its Place. Chicago: University of Chicago Press, 2003. Mannheim, Karl. Ideology and Utopia. New York: Harcourt Brace Jovanovich, 1936. Merton, Robert K. Social Theory and Social Structure. New York: Free Press, 1968. Motherwell, Robert et al., eds. Modern Artists in America. First Series. New York: Wittenborn Schultz, 1951. Nagel, Thomas. The View from Nowhere. Oxford: Oxford University Press, 1986. Palmer, Vivien M. Field Studies in Sociology. A Student’s Manual. Chicago: University of Chicago Press, 1928. Park, Robert E. and Ernest W. Burgess. The City. Chicago: University of Chicago Press, 1925. Persons, Stow. Ethnic Studies at Chicago, 1905-45. Urbana: University of Illinois Press, 1987. Pickering, Andrew. The Mangle of Practice. Time, Agency, and Science. Chicago: University of Chicago Press, 1995. Rosenberg, Bernard and Norris Fliegel. The Vanguard Artist. Chicago: Quadrangle Books, 1965. Rosenberg, Harold. “Gottlieb.” Adolph Gottlieb. Paintings 1950-1971. London: Marlborough Gallery, 1971. Rosenberg, Harold. “Tenth Street. A Geography of Modern Art.” Discovering the Present. Chicago: University of Chicago Press, 1973. 100-09. Rosenberg, Harold. Willem deKooning. New York: Harry N. Abrams, 1973. Sassen, Saskia. The Global City. 2nd ed. Princeton: Princeton University Press, 2001. Sawin, Martica. Surrealism in Exile and the Beginning of the New York School. Cambridge: MIT Press, 1995. Seidner, David. Artists at Work. Inside the Studios of Today’s Most Celebrated Artists. New York: Rizzoli, 1999. Seitz, William C. Abstract Expressionist Painting in America. Cambridge: Harvard University Press, 1983. Seuphor, Michel. “Paris New York 1950.” Modern Artists in America. First Series. Ed. Robert Motherwell et al. New York: Wittenborn Schultz, 1951. 118-22. Shaw, Clifford R. The Jack-Roller. A Delinquent Boy’s Own Story. Chicago: University of Chicago Press, 1930.
420
Thomas F. Gieryn
Smith, T.V. and Leonard D. White, eds. Chicago. An Experiment in Social Science Research. Chicago: University of Chicago Press, 1929. Thistlewood, David. “Historicity and Mythology in American Abstract Expressionism.” American Abstract Expressionism. Ed. idem. Liverpool: Liverpool University Press, 1993. 1-39. Thrasher, Frederic M. The Gang. A Study of 1313 Gangs in Chicago. Chicago: University of Chicago Press, 1927. Wirth, Louis. The Ghetto. Chicago: University of Chicago Press, 1928. Wolfe, Tom. The Painted Word. New York: Bantam, 1975. Zorbaugh, Harvey W. The Gold Coast and the Slum. Chicago: University of Chicago Press, 1929.
GEORGES DIDI-HUBERMAN
The Eye Opens, the Lamp Goes Out: Remarks on Bergson and Cinematography At the very time when the experimental method theorized by Claude Bernard was triumphing in the biology laboratories, Henri Bergson proposed that we “seek experience at its source, or rather above that decisive turn where, taking a bias . . . it becomes properly human experience.”1 Not that Bergson rejected Bernard’s teaching. On the contrary, he praised his modesty in the face of the real, his way of respecting the singularity of phenomena, that is to say their multiplicity and complexity. In the same way, he admired the enormous philosophical daring of rejecting all spirit of system in the name of experience (Bergson quotes this dictum by Bernard: “Philosophy and science should not be systematic.”) to produce authentic experimental concepts, which would be fluid concepts or at least flexible, plastic. But let us also remember that an idea no matter how flexible we may have made it, will never have the same flexibility as a thing. Let us therefore be ready to abandon it for another, which will fit the experiment still more closely.2
Philosophy, says Bergson elsewhere, is strictly itself only when it goes beyond the concept, or at least when it frees itself of the inflexible and ready-made concepts and creates others very different from those we usually handle, I mean flexible, mobile, almost fluid representations, always ready to mold themselves on the fleeting forms
of a real, which analysis alone remains incapable of apprehending in its singular movements.3 But what is a thinking of the real that is capable of renouncing not only the systems in general, but also the rigidity – if not the rigor – of the concepts themselves? 1 2 3
Henri Bergson. Matter and Memory. Trans. Nancy Margaret Paul and W. Scott Palmer. New York: Zone Books, 2002. 184. Bergson. The Creative Mind. Trans. Mabelle L. Andison. New York: Greenwood Press, 1968. 245f. Ibid. 168.
422
Georges Didi-Huberman
Each reader of Bergson remembers the response to this question as a particularly shattering experience of thought, when at the beginning of Matter and Memory the young philosopher demands that we place ourselves in a situation – typically experimental – where “we will assume for the moment that we know nothing,” thus renouncing all anchorings in existing philosophical systems – either realistic or idealistic – in order to “hold” at first “to the appearances.” What happens then? “Here I am in the presence of images, in the vaguest sense of the word, images perceived when my senses are opened to them, unperceived when they are closed” (italics added).4 It is easy to imagine, on hearing such words, all the great founders of systems, from Plato to August Comte, furiously turning in their graves. What kind of philosophy demands that appearance should be maintained to better think the appearing, and, in its flux, the structure of the real itself? What kind of philosophy dares to rely on images in their “vaguest sense,” and which does not fear, a few pages on, to assert: “the material world is made of objects, or, if you prefer it, of images.” And even: “I call matter the aggregate of the images.”5 What kind of philosophy allows perception and memory to interpenetrate, as far as demanding the notion of “unconscious representation”?6 The audacity of this entry into philosophical matter in the first chapter of Matter and Memory has nothing of that metaphysical or “vitalist” urge in which one all too easily encloses Bergson’s thought. This audacity, in effect, produces the rigorous counterpoint to a veritable epistemology in formation, which is to say a reflection on the stakes and limits of positive science. Well before Gaston Bachelard, Bergson posed the problem of scientific procedure in terms of obstacles and “false problems”; well before Louis Althusser, he underlined the misdeeds of a “spontaneous philosophy of scholars” by opposing the poverty of scientism to the richness of science itself.7 The philosopher must also rethink what experience can mean in the triumphant age of experimental method, which supposes a new relationship between thought and knowledge: At first sight, it may seem more prudent to leave the consideration of facts to positive science, to let physics and chemistry busy themselves with matter, the biological and psychological sciences with life. The task of the philosopher is then clearly defined. He takes facts and laws from the scientists’ hand; and 4 5 6 7
Bergson. Matter and Memory. 17. Ibid. 22, 68. Ibid. 65, 133 passim. Bergson. Creative Mind. 33-107.
The Eye Opens, the Lamp Goes Out
423
whether he tries to go beyond them in order to reach their deeper cause, or whether he thinks it impossible to go further and even proves it by the analysis of scientific knowledge, in both cases he has for the facts and relations handed over by science, the sort of respect that is due to a final verdict. To this knowledge he adds a critique of the faculty of knowing, and also, if he thinks proper, a metaphysic; but the matter of knowledge he regards as the affair of science and not of philosophy. But how does he fail to see that the real result of this so-called division of labor is to mix up everything and confuse everything? The metaphysic or the critique that the philosopher has reserved for himself he has to receive, ready-made, from positive science, it being already contained in the descriptions and the analyses, the whole care of which he left to the scientists. For not having wished to intervene, at the beginning, in questions of fact, he finds himself reduced, in questions of principle, to formulating purely and simply in more precise terms the unconscious and consequently inconsistent metaphysic and critique which the very attitude of science to reality marks out.8
And Bergson warns us (as a first justification of his theoretical audacities): “Philosophy, then, invades the domain of experience . . . At first there may be a certain confusion.”9 What does this mean? That philosophy, in this domain, only achieves its precision – these are the stakes claimed on the first pages of La Pensée et le Mouvant10– by exploding the normal frame of its relationship to science, in creating the inevitable confusion of a displacement of the limits between the established methods and domains of knowledge. The idea that philosophy is something like a “synthesis of the positive sciences” expressed in normal language, proved itself, in Bergson’s eyes, as just as ‘disobliging for science’ (not needing to draw conclusions for it from its own procedure) as it is ‘injurious to philosophy’ (not needing to say in its place what is real and what is not).11 Where, exactly, does this necessary confusion come from, which philosophical work should be capable of transforming into an instrument of precision? It comes as a result of the fact that Bergson does not hesitate to face up to the real – matter, life, movement, duration, consciousness, memory – through the image of its appearance, the image in so far as it is not reduced to a simple perception. Such is the chief philosophical decision,12 a decision that entails a methodological overthrowing 8 9 10 11 12
Bergson. Creative Evolution. Trans. Arthur Mitchell. New York: Dover, 1998. 194. Ibid. 198. Bergson. Creative Mind. 9-33. Ibid. 145. Recognized by a few exegetes of Bergson’s thought (cf. especially Lydie Adolphe. La dialectique des images chez Bergson. Paris: P.U.F, 1951; Paul Naulin. “Le pro-
424
Georges Didi-Huberman
of knowledge itself. A stable thing (already appeared and not yet disappeared, as if fixed in being) is susceptible to analysis because one can move around it producing a regulated series of points of view of the subject, whereas the image of appearance (the non-fixable appearance, the appearing thing) is only susceptible to intuition, a movement of the mind forcing us to ‘enter’ the image if we do not want to miss the thing completely.13 To know (connaître) through images therefore means to approach the appearance of things from this side of the observable fact. It means to touch the singularity on this side of all generalizing laws. For “one blade of grass does not resemble another blade of grass any more than a Raphael resembles a Rembrandt.”14 In writing this, Bergson naturally does not think of casting the normal botanical classifications into doubt; but he demands for every “blade” of being, if I may use this term, that the particular style of its appearing is recognized in a constantly different temporality and context. That is why knowing through images situates itself on this side of the representation, and the impoverishment which this imposes (through abstraction, through schematization, through “prevision”) on the unforeseeable movement of appearance.15 A botanist can quite simply, when he classifies, announce the identity of two blades of grass; the Bergsonian philosopher – in this sense closer to an art historian – would prefer to consider their singularity, their similarities and dissimilarities.16 To know through images means to do without the synthesis of the “already-made” (tout fait) and to accept the risk of the – inevitably provisional but rhythmicized by time in actu – intuition of the “being-made”
13 14 15 16
blème de la conscience et la notion d’‘image.’” Bergson. Naissance d’une philosophie. Actes du colloque de Clermont-Ferrand. Paris: P.U.F, 1990. 97-109; JeanFrançois Bordron. “Bergson et les images. L’iconicité de la pensée dans ‘Le possible et le réel.’” Lire Bergson. Le possible et le réel. Ed. Frédéric Cossutta. Paris: P.U.F, 1998. 159-81) and developed above all in Deleuze’s philosophy of the image (cf. especially Gilles Deleuze. The Movement-Image. Cinema 1. Trans. Hugh Tomlinson and Barbara Habberjam. London: Athlone Press, 1986; and. The Time-Image. Cinema 2. Trans. Hugh Tomlinson and Robert Galeta. London: Athlone Press, 1989). But whatever Deleuze may say about it, the Bergsonian attempt is in many points related to the phenomenological approach – if not reduction – of appearance. Bergson. Creative Mind. 159-62. Ibid. 122. Ibid. 107. “I believe that identity is something geometrical and resemblance something vital. The first has to do with measure; the other belongs rather to the domain of art.” Ibid. 67.
The Eye Opens, the Lamp Goes Out
425
(se faisant).17 It means to achieve that which in each system “is worth more than the system and survives it.”18 It means to discover that beyond the arrangements or “juxtapositions” which the real would be mechanically and intemporally composed of, we must take account of a perpetual creation, i.e. an unforeseeable rearrangement of all things, which proceeds by the interpenetration of all these things among themselves, in space as in time.19 From here one comes to the fruitful paradox of a practice of thought defined as a “faculty of seeing which is immanent” in movement and duration, a practice where the “understanding itself, by submitting to a certain discipline, might prepare a philosophy which transcends it,”20 that is to say a philosophy capable of eluding the petrifaction of the system, even the rigidity of the concept. When Bergson calls for the “adhesion” of the concept to its object – against “conceptions so abstract, and consequently so vast, that it might contain, aside from the real, all that is possible and even impossible”21 – he clearly signifies, without fear of paradoxes, that a concept worthy of the name must be a singular instrument, “moldable” to its single object, and consequently resistant to every generalization, though capable of the fluidity and plasticity which plaster or modeling clay demonstrate on the constantly changing mold into which they are pressed.22 To know through images means therefore to touch the real precisely by means of the detour, the power of immanence of which images are the privileged vehicle: the molding of things (as the plaster works of Rodin, in Bergson’s time, retained the singularity of the least element of his formal vocabulary) and the modulation of the milieu (as Claude Monet’s palette, which in the same period reconfigured the fluidity and the nuances of the smallest drop of water in the Water Lilies). But what the image gains in precision and nuances, in singularity and multiplicities, in “molding” and “modulations,” it loses in universality (since the molding only adapts to the unique form of which it offers the negative) and stability (because a modulation never stops changing). It is a case here of an uncertainty principle avant la lettre: 17 18 19 20 21 22
Bergson. Creative Evolution. 237. Ibid. 238. Ibid. Ibid. XV and 250. Bergson. Creative Mind. 9. Ibid. 150. One must recall that Bergson contrasts, on this point, the fluidity of the analytical and experimental concepts of Aristotle with the rigidity of the systematic and ideal concepts of Plato. Cf. Ibid. 261-300.
426
Georges Didi-Huberman
the same phenomenon cannot be simultaneously observed in its two principle modes of existence, corpuscular (mechanical) and undular (dynamic). The image, because it molds itself on the singularity of phenomena, can restore the “thousand shades;”23 but precisely its adhesive power, its capacity of immanence, prevents it from giving us a continual idea, permanent or eternal, of the real (and precisely that is the reason why, according to Bergson, it is necessary entirely to rethink the connections between science and metaphysics).24 Image and intuition, which fail where concept and synthesis show themselves effective, succeed where concept and synthesis stumble; they “extend our perception” by respecting the differences, the shades, the movements, and the smallest qualitative changes in reality.25 But there are at least two compensations: the thing disappears in its stability, in its quantity or measure, its nature as an “immobile section,” in favor of a continuous duration, which forms, in a certain way, the moving milieu of appearance; the time of cognition becomes a discontinuous time, unstable, intuitive, jerky, fading because “molded” on the singular time of appearance. To think intuitively is to think in duration. Intelligence starts ordinarily from the immobile, and reconstructs movement as best it can with immobilities in juxtaposition. Intuition starts from movement, posits it, or rather perceives it as reality itself, and sees in immobility only an abstract moment, a snapshot taken by our mind, of a mobility. Intelligence ordinarily concerns itself with things, meaning by that, with the static, and makes of change an accident which is supposedly superadded. For intuition the essential is change: as for the thing, as intelligence understands it, it is a cutting which has been made out of the becoming and set up by our mind as a substitute for the whole . . . Intuition, bound up to a duration which is growth, perceives in it an uninterrupted continuity of unforeseeable novelty.26 . . . Intuition is there, however, but vague and above all discontinuous. It is a lamp almost extinguished, which only glimmers now and then, for a few moments at most . . . These fleeting intuitions, which light up their object only at distant intervals, philosophy ought to seize, first to sustain them, then to expand them and so unite them together.27
We have before us, in a strict sense, only moving things: the world is the moving. But how should one recognize (connaître) the movements of the moving? It seems that here Bergson confronts us with a contradiction: on the one hand one should resist thinking of movement in discontinuous terms, stop reducing movement to “snap shots” or “juxta23 24 25 26 27
Ibid. 164. Ibid. 188. Ibid. 153-87. Ibid. 38f. Bergson. Creative Evolution. 267-69.
The Eye Opens, the Lamp Goes Out
427
posed immobilities;” on the other hand the seizure of movement – intuition, image – only happens in the mode of the “vague” and “above all through the discontinuous.” Our thought, writes Bergson, only lights up the phenomenon like “a lamp almost extinguished, which only glimmers now and then, for a few moments at most.” The intuition captures the moving in as much as it – since it is immanent to the moving – passes like a butterfly appearing and fading almost immediately in the opaque sky of human intelligence. The image “extends perception” of things, but the price to pay lies in the singularity, therefore the fragility and even the “transience” – to speak with Freud – of the image.28 It is not by chance that Bergson demands that we “plunge [into the perception of phenomena] for the purpose of deepening and widening it,” in the manner of Turner’s paintings when they justly capture in untiring variations the “dissolving views” of wind, fog, or storms.29 Bergson’s paradigm of “a lamp almost extinguished” would therefore be – as was anticipated in the Essai sur les données immédiates de la conscience with the question of phenomenon envisaged from the angle of its intensity – an aisthetic or even aesthetic paradigm.30 It is not surprising, let it be said, to see this paradigm re-emerge at crucial moments of an art discourse, when it is the case precisely of affirming that something new is in the process of appearing. In this way Jean Paulhan evokes “a small nocturnal adventure” – to light and immediately extinguish a lamp in order to feel about in the space fleetingly seen of an apartment – in order to show the confusing value of that which Cubist space allows to be seen.31 Similarly, Tony Smith later relates his need to sculpt minimalist volumes to the sensory experience of a nocturnal journey suddenly deprived of its normal lighting, on a highway under construction in New Jersey.32 Earlier, Edmund Husserl had tried to forge a concept of “empty representation” by questioning, in the same example of “the light that one 28
29 30
31 32
That is why Bergson demands the image only to multiply it: images of “flowing,” then “the unrolling of a spool,” or a ball of thread, then of a “spectrum of a thousand shades,” then of an “infinitely small piece of elastic,” etc. Bergson. Creative Mind. 162-65. Ibid. 158, 160. Henri Bergson. Time and Free Will. An Essay on the Immediate Data of Consciousness. Trans. F.L. Pogson. London: George Allen and Company, 1912. 1-5. Here the question concerning intensity is related from the beginning to the example of “deep feelings” and “aesthetic feelings.” Jean Paulhan. La peinture cubiste (1942-1956). Paris: Denoël, 1990. 61. On Tony Smith, cf. Georges Didi-Huberman. Ce que nous voyons, ce qui nous regarde. Paris: Éditions de Minuit, 1992. 61.
428
Georges Didi-Huberman
extinguishes,” what one still sees (as image) in that which has just disappeared (as phenomenon).33 But it is another reality – equally philosophical – to which the Bergsonian configuration refers us still more directly. An instrumental dipositif made of images that appear in order to disappear immediately, it aims at the continuity of movement, but in order to render this visible, uses the artifice of a stroboscopic intermittency. This dispositif, which came about in Bergson’s time, offered the singular experience of a visual regime defined by the jerky character of the passing images and the luminous intensity, as if a lamp obstinately extinguishes itself and relights itself on the visible world with every turn of the projectionist’s crank. This experience is that of cinematography. In his introductory chapter to the Movement-Image, Gilles Deleuze in a certain way “rescued” the Bergsonian critique of the “cinematographic illusion” by showing that it allows us, in the most general frame of a philosophy of movement and time, to make the cinema a little more thinkable.34 In fact, Bergson seems to have refuted the cinema as someone who holds a piece of film in their hands where movement is reduced to the “immobile sections” of a photogram, as opposed to a projected film – what Deleuze names, in contrast, “mobile sections.” And Deleuze writes, to raise the stakes: “Is not the reproduction of the illusion in a certain sense also its correction? Can we conclude that the result is artificial because the means are artificial?”35 In short, film as support constitutes, in Bergsonian terms, a lie about movement; the projected film can offer the best means to rediscover the truth of movement as the “immediate given.”36 Deleuze even remarks that only the “modern” cinema – since Roberto Rosselini – would be capable of embodying this time-image, for which, in his eyes, Bergson’s thought calls.37 Nevertheless, it is necessary, it seems to me, to return to the conditions themselves under which Bergson contrasted his own notion of movement-image or duration-image with what he called a cinematographic illusion. What exactly does he aim at with the word “cinematography”? In the year 1907, the reader of Creative Evolution, which 33
34 35 36 37
Edmund Husserl. Logische Untersuchungen. Ed. Ullrich Melle. Dordrecht, Boston, and London: Kluwer Academic Publishers, 2002, supplementary volume, part one (= Husserliana, vol. XX/1). 141. Deleuze. Movement-Image. 1. Ibid. 2. Ibid. Deleuze. Time Image. 62, 354 passim.
The Eye Opens, the Lamp Goes Out
429
included the famous chapter “The Cinematographical Mechanism of Thought and the Mechanistic illusion,”38 could already see a considerable number of short films by the Lumière brothers – some thousand films were already in their catalogue of 1905 – and no less than fiftyfour films by Georges Méliès – including Le Rêveur éveillé, La Toile d’araignée merveilleuse, Hallucinations pharmaceutiques, or also La Poupée vivante – were shown in 1907 at the Théâtre Robert-Houdin.39 But Bergson wanted to make clear that his “cinematographic” analogy went back to a lecture at the Collège de France in 1902/03. “We then compared the mechanism of conceptual thought to that of the cinematograph,” he remarked.40 Intelligence as artificial instrumentalization of duration, mechanistic thought as reduction of movement to simple sections of instants or to spatial arrangements, all these themes in Creative Evolution41 are not new in Bergson, far from it. They already existed at a time when the Lumière brothers were still a long way from patenting their invention of the “Cinematograph.” One finds them not only in 1896 in Matter and Memory,42 but already in 1889 in Essai sur les données immédiates de la conscience, where Bergson analyzed the two antagonistic concepts of duration by criticizing the mechanistic reduction of intensive phenomena to simple problems of extension, measure, or divisibility.43 On the other hand, one is surprised that no allusion to the comic cinema is made in Le Rire in 1900, while numerous comical views by the Lumière brothers and cinematographic farces by Méliès – with charming titles like Spiritisme abracadbrant, Le Déshabillage impossible or Le Savant et le chimpanzé, to take examples from this year alone – would reveal the complete effectiveness of the “something mechanical encrusted on the living.”44 When, later, Bergson continues to criticize the “simple snapshots which our understanding has taken of the continuity of movement and duration,” when he castigates the “succession similar to that of the images of a cinematographic film,” which our perception 38 39
40 41 42 43 44
Bergson. Creative Evolution. 272. Cf. Bernard Chardère. Les images des Lumière. Paris: Gallimard, 1995. 91. Jacques Malthête and Laurent Mannoni, eds. Méliès. Magie et cinéma. Paris: Éditions Paris-Musées, 2002. 261. Bergson. Creative Evolution. 272. Ibid. Bergson. Matter and Memory. 324-29, 337-43. Bergson. Immediate Data. xix, 1, 5, 69, 90 and passim. Bergson. “Laughter”. Comedy. Ed. Wylie Sypher. Baltimore: Johns Hopkins University Press, 1956. 84 passim.
430
Georges Didi-Huberman
remains “forced to divide up . . . image by image,”45 we must understand, once and for all, that Bergson’s critique, however aesthetic at its base, was in no way directed against the “seventh art” of cinematography. Indeed, for Bergson, the cinematographic image was related to the moving image, as all mechanistic cognition (abstract, external) to dynamic cognition (concrete, inherent) of phenomena. “Cinematography,” as Bergson understood it, derives from a gnoseological paradigm, and not the burlesque theaters of the grands boulevards at the end of the nineteenth century equipped with the invention of Edison and Lumière. Because it is always mentioned in the context of a philosophy of cognition and sensation, Bergsonian “cinematography” should be understood in the strict sense attributed to it by Lucien Bull in 1928 in his book La Cinématographie. It is the case of an experimental instrument of visual cognition based on the analytical dissection of movement, and aiming at its theoretical synthesis.46 Lucien Bull was the last assistant of Étienne-Jules Marey, and in addition, his “spiritual son.”47 It is therefore in the context of chronophotography and Marey himself that we must place what Bergson understood by cinematography. When the philosopher entered the Collège de France, the great physiologist had already been teaching there for around thirty years. Though they were probably not made to properly understand each other, they could not ignore each other either. One occasionally has the impression that La pensée et le mouvant was written in response to Marey’s Le mouvement, that Creative Evolution inverts all propositions in La machine animale, and that Matter and Memory tries to entirely refute La methode graphique – this mnemotechnique of phenomena – demanded by the learned mechanicalist.48 Nevertheless, Bergson and Marey were members of the same academic societies, and they co-signed a program for experimental studies of certain “psychophysical” phenomena linked to hypnosis and the nineteenth century spiritualist vogue.49 45 46 47 48
49
Bergson. Creative Mind. 15-18. Lucien Bull. La cinématographie. Paris: Armand Collin, 1928. vii, 1. Cf. Laurent Mannoni. Étienne-Jules Marey. La mémoire de l’œil. Paris and Milan: Mazotta/La Cinémathèque Française, 1999. 383. Cf. Étienne-Jules Marey. La machine animale. Paris, 1873; idem. La méthode graphique dans les sciences expérimentales et particulièrement en physiologie et en médecine. Paris, 1878; idem. Le mouvement. Paris, 1894. Henri Bergson, Étienne-Jules Marey et al. “Groupe d’études de phénomènes psychiques (1901)”. Mélanges. Ed. André Robinet. Paris: P.U.F, 1972. 509.
The Eye Opens, the Lamp Goes Out
431
Marey is never quoted in Bergson’s books, but allusions to his work can undoubtedly be found there. When, in the Essai sur les données immédiates de la conscience, Bergson asserts that movement is as little divisible as duration is measurable, then Marey’s whole attempt – with his visual fragmentation of gestures and the concomitant need to measure – is philosophically put in question; when Bergson strongly criticized those “who are content to juxtapose states [and] form from this a chain or line,” he seems to reject Marey’s chronophotographic series as the countless curves which were meant to give a legible trace – simultaneously indicatory and geometric – of vital phenomena (fig. 1 and 2).50 Certain theoretical comparisons in Matter and Memory give the impression that Bergson thought them up because he had Marey’s plates – produced at the same time – in front of him. Therefore, the “thousand successive positions of a runner” create a common object of inquiry for philosopher and physiologist alike (fig. 3); but here Bergson demands the “change in depth” (or what he had already termed the “inner multiplicities”) in opposition to the change that we “localize here and there, but on the surface” which he probably saw in the records of Marey’s experiments.51 On other pages of the same book, Bergson refutes everything which serves Marey to capture the dynamic of the élan vital, to measure and synthesize. “The axes or points, to which one sets it in relation,” express just as little of the movement of a body as the juxtaposed elements of a chronographic series; the line – “divisible [and] deprived of quality” – describes no more than “mathematical symbols,” because in both cases it is a matter of “restricting movement in space,” by ignoring its real temporality, its own rhythmicity. Bergson even recalls “the black screen on which the image could be shown,” exactly as one sees it in the photographic procedures that Marey uses in his experimental “station” (fig. 4).52 It is therefore consistent that in Creative Evolution the “cinematographic illusion” is described in terms – typical for Marey – of “snapshots,” which pretend to capture movement by juxtaposing in space a series of “immobile sections:” The intellect, like the senses, is limited to taking, at intervals, views that are instantaneous, and by that very fact immobile, of the becoming matter . . . Thus, we pluck out of duration those moments that interest us, and that we have gathered along its course. These alone we retain [and] we become unable to perceive the true evolution, the radical becoming. Of becoming we perceive only 50 51 52
Cf. Bergson. Time and Free Will. 69. Marey. Méthode. 1ff. Bergson. Matter and Memory. 207. Ibid. 39.
432
Georges Didi-Huberman
Fig. 1: Étienne-Jules Marey. Man walking (wearing a black suit with white stripes) (1884).
Fig. 2: Étienne-Jules Marey. Horse in step: geometrical sketch (1885).
The Eye Opens, the Lamp Goes Out
433
Fig. 3: Étienne-Jules Marey. Man walking (wearing a black suit with a white stick fixed along the spinal column) (1886).
Fig. 4: Étienne-Jules Marey. Knee bends, arms extended (man wearing a black suite with white stripes) (1884).
434
Georges Didi-Huberman
states, of duration only instants, and even when we speak of duration and of becoming, it is of another thing that we are thinking. Such is the most striking of the two illusions we wish to examine. It consists in supposing that we can think the unstable, the moving by means of the immobile . . . Now, life is an evolution. We concentrate a period of this evolution in a stable view which we call a form, and, when the change has become considerable enough to overcome the fortunate inertia of our perception, we say that the body has changed its form. But in reality the body is changing form at every moment; or rather, there is no form, since form is immobile and the reality is movement. What is real is the continual change of form: form is only a snapshot view of a transition. Therefore, here again, our perception manages to solidify into discontinuous images the fluid continuity of the real.53
That this illusion is as effective as a conjuring trick – in the style of Méliès or Robert-Houdin – is shown by the “artifice of cinematography” that Bergson once described in its extravagant dimension of a film projected onto a screen: a trick because it only gives back from a singular movement a sketch of “movement in general” integrated into the mechanical apparatus.54 This analysis does not only neglect the concrete cinematographic experience – because it denies the indexicality, i.e. the ability to give back singular movements – but what is more, it leads Bergson to propose later the absurd argument according to which “the film could be run off ten, a hundred, even a thousand times faster, without the slightest modification in what is being shown.”55 This relative lack of attention to the cinema as a specific sensory experience shows, if that is necessary, that the problem for Bergson lies elsewhere. In Creative Evolution, it is the century-old question about form and time which forms the core of his critical work. “Form is only a snapshot taken of a transition:” in this sentence the word “snapshot” denotes a recent photographic invention – the development of “extra rapid” silver gelatine plates – from which Marey drew the total effectiveness of his instrumental method.56 The word “form” on the other hand refers to a long philosophical tradition that goes back to at least Plato. It is precisely the form as eidos or idea, as metaphysical abstraction, from which Bergson (almost a century before Jacques Derrida) departs. This is why it is not out of place, in his eyes, to use other much 53 54 55
56
Bergson. Creative Evolution. 272, 302. Ibid. 305. Bergson. Creative Mind. 18. This argument corresponds, incidentally, to the one that Bergson raises in relation to mechanistic cognition: the laws of physics would not change if the world moved ten times faster or slower, as it does in reality, a proof that these laws lack the concrete element of duration. Cf. Michel Frizot. Étienne-Jules Marey. Chronophotographie. Paris: Nathan/ Delpire, 2001. 103.
The Eye Opens, the Lamp Goes Out
435
less “modern” comparisons than that of cinematography: the kaleidoscope, where each configuration (form) should not allow us to forget the jolt (time) which engendered it; the mosaic, where the reunion of all the elements (the divided labor of form) never succeeds in reproducing the simple stroke of the drawing of the master (the indivisible work of time), which served as its model; up to the bas-reliefs of the Parthenon, where Bergson sees “the same cinematographic mechanism” at work, if it is only that antique metaphysics isolates “characteristic poses” where modern science works with “any instant whatever,” which they “place all in the same class” in the same law of physics or on the same chronophotographic plate.57 One understands, therefore, to what extent “cinematography” describes less a specific instrument than a very general instrumentalization, where appearance finds itself, in a certain way, caught and denied. Marey’s chronophotography is only the contemporary example of a mythology of knowledge that has existed since the Eleatic philosophers wanted to reduce movement to the space traversed by the moving body.58 Marey represents for Bergson the modern incarnation of an Argus who never wants the lamp to go out nor the eye to close: the “cinematographic illusion” which consists of believing one sees everything and of manipulating this everything seen like an extension which is divisible at will, a quantity integrally geometrizable, an easily instrumentalizable vital energy. Now, we know from experience that appearance, including its visual dimension, is vital duration: that is to say, it does not endure as something hardened, stiff, so-called stable and susceptible, which is suited to give rise to some eternal idea; but rather that it endures as the dancing, untiring flux of the waves. More profoundly, appearance endures, in Bergson’s sense, to the extent that it passes by, by surviving in a movement image.59 Further, it is necessary to know – and this would be the object of an “experimental philosophy” as proposed by Bergson – how to draw out an inherent knowledge from this passage, this survival, this movement or trail (sillage) of appearance . . .60 Translation: Benjamin Carter 57 58 59 60
Bergson. Creative Evolution. 303, 332. Bergson. Time and Free Will. 75. Bergson. Matter and Memory. 133. (“Of The Survival of Images. Memory and Mind.”) I develop this notion of “sillage” in an article that takes its support from this one and extends it. Georges Didi-Huberman. “L’image-sillage.” L’Inactuel 10 (2004).
436
Georges Didi-Huberman
WORKS CITED Adolphe, Lydie. La dialectique des images chez Bergson. Paris: P.U.F, 1951. Bergson, Henri. Time and Free Will. An Essay on the Immediate Data of Consciousness. Trans. F.L. Pogson. London: George Allen and Company, 1912. Bergson, Henri. “Laughter.” Comedy. Ed. Wylie Sypher. Baltimore: Johns Hopkins University Press, 1956. Bergson, Henri. The Creative Mind. Trans. Mabelle L. Andison. New York: Greenwood Press, 1968. Bergson, Henri, Étienne-Jules Marey et al. “Groupe d’études de phénomènes psychiques (1901).” Mélanges. Ed. André Robinet. Paris: P.U.F, 1972. Bergson, Henri. Creative Evolution. Trans. Arthur Mitchell. New York: Dover, 1998. Bergson, Henri. Matter and Memory. Trans. Nancy Margaret Paul and W. Scott Palmer. New York: Zone Books, 2002. Bordron, Jean-François. “Bergson et les images. L’iconicité de la pensée dans ‘Le possible et le réel.’” Lire Bergson. Le possible et le réel. Ed. Frédéric Cossutta. Paris: P.U.F, 1998. 159-81. Bull, Lucien. La cinématographie. Paris: Armand Collin, 1928. Chardère, Bernard. Les images des Lumière. Paris: Gallimard, 1995. Deleuze, Gilles. The Movement-Image. Cinema 1. Trans. Hugh Tomlinson and Barbara Habberjam. London: Athlone Press, 1986. Deleuze, Gilles. The Time-Image. Cinema 2. Trans. Hugh Tomlinson and Robert Galeta. London: Athlone Press, 1989. Didi-Huberman, Georges. Ce que nous voyons, ce qui nous regarde. Paris: Éditions de Minuit, 1992. Didi-Huberman, Georges. “L’image-sillage.” L’Inactuel 10 (2004): 111-126. Frizot, Michel. Étienne-Jules Marey. Chronophotographie. Paris: Nathan/Delpire, 2001. Husserl, Edmund. Logische Untersuchungen. Ed. Ullrich Melle. Dordrecht, Boston, and London: Kluwer Academic Publishers, 2002, supplementary volume, part one (= Husserliana, vol. XX/1). Malthête, Jacques and Laurent Mannoni, eds. Méliès. Magie et cinéma. Paris: Éditions Paris-Musées, 2002. Mannoni, Laurent. Étienne-Jules Marey. La mémoire de l’œil. Paris and Milan: Mazotta/ La Cinémathèque Française, 1999. Marey, Étienne-Jules. La machine animale. Paris, 1873. Marey, Étienne-Jules. La méthode graphique dans les sciences expérimentales et particulièrement en physiologie et en médecine. Paris, 1878. Marey, Étienne-Jules. Le mouvement. Paris, 1894. Naulin, Paul. “Le problème de la conscience et la notion d’‘image.’” Bergson. Naissance d’une philosophie. Actes du colloque de Clermont-Ferrand. Paris: P.U.F, 1990. 97109. Paulhan, Jean. La peinture cubiste (1942-1956). Paris: Denoël, 1990.
MARTIN BURCKHARDT
The Illusion of Power: Central Bank Money I. Ladies and Gentlemen,1 before turning my scrutiny upon a rather curious instrument, namely the banknote [Schein],2 let me say a few words about the historical issues which the seventeenth century raises for us. As we know (miraculously prompted from the wings by Schlegel), the historian is a prophet in reverse. That is to say, he does not approach his subject gingerly, but rather with a prophetic interest – the same approach my father took to our Burckhardt genealogy. Please forgive this private detail; it is quite illuminating. With an unerring sense of favorable lines, with a feel for marital politics in reverse, my father tracked down the noble strands in the family tree – and to miraculous effect: he was able to trace back the clan’s genealogy not to some modest Burckhardt bourgeoisie, but to 1
2
This essay, originally written as a lecture, follows the rules of spoken language, oral history. Rather than camouflage this with a thin veneer of textualization, the direct address remains – one ought perhaps to fill in the absent audience, the people who assembled early one Saturday morning to hear about the seventeenth century, only to be swept off to times still more remote. Of course, there is a deeper reason for this fidelity to the origins: the lecture form, the dialogue with a silent audience, allows for intellectual short-cuts, rhetorical maneuvers which make it possible to condense relatively complex connections as if in fast-forward. In this spirit, the essay summarizes thoughts which have been worked out with much greater precision elsewhere – and does so with a brevity which, though it seems almost inadmissible in the world of books, is perfectly tailored to a discursive culture that has mislaid the perseverance of paper. Which in a certain respect, and somewhat ironically, affects the manuscript of this lecture as well, as the most important passages were not read that Saturday morning – due to time constraints. The German word for “banknote” is Schein, which also means “illusion,” “appearance,” and “shine” or “radiance” – an ambiguity that is crucial to Burckhardt’s argument. In the contexts that maintain and indeed depend on this ambiguity, I have left the word in the original [translator’s comment].
438
Martin Burckhardt
royal houses, with Charlemagne right at the top (which profoundly impressed me as a child, standing before this illustrious family tree). My father may have been slightly over the top in this regard, but the tendency is telling – one is always in danger of appropriating history as a prehistory for Now. In this spirit – paraphrasing not my father now, but Winston Churchill – my first impulse would be: “Never trust a story you didn’t fake yourself.” Which could also be formulated positively, in the form of a question, or more precisely, in a question as to the question: What might we want from the seventeenth century? In what way and with what ulterior motives do we pick out a bloodline? Which especially flattering bloodlines are accentuated? How is it that the seventeenth century appears to us as a kind of rearview mirror of rationality? At any rate – one hardly needs to explain that this is the case. Let us take the most momentous figure of thought that period bequeathed us – the Leviathan – and then look at the world map. What do we see? We see that there is no longer a spot left on the globe that is not organized as a state. Even if many of these entities, created on the drawing board, still have a tribalistic structure, even if the concept of the state has really taken root and become institutionalized only in a small part of the world, the fact remains that the form of the territorial state has become a universal. Indeed, it is no wonder that the seventeenth century, which formed states, clothed its soldiers in uniforms, taxed its subjects and finally, at the end of the century, instituted the first central bank, the Bank of England, strikes us as the perfect epoch for reflecting us and our society. Of course one is reluctant to emphasize this aspect when it sounds much nobler to speak of the birth of the modern subject or the like – very much in my father’s spirit. After all, this pulls off the fine trick of making the subject’s newly acquired subjection into a kind of title of nobility. I think you can tell what I’m getting at: the seventeenth century, if I may put it that way, is to scholarship what Charlemagne is to my family tree. That sounds awfully flippant. Yet I haven’t made it easy for myself, considering that in my reflections on machines I have kept stumbling across the seventeenth century from the very start. If I speak of stumbling, it is because this time, this epochè, strikes me as a discursive border between day and night – as if all that lies before were swallowed by darkness. In that respect, what I have referred to as a rearview mirror is a retroactive process: an obfuscation after the fact. One quick example (because we are discussing instruments here): We know that the mechanical clock is the central stratagem of the Cartesians. But the mechanical clock itself is a product of the twelfth or thirteenth century. If
The Illusion of Power: Central Bank Money
439
we take the machine as the engine of thought and philosophy – as René Descartes incessantly encourages us to – we are faced with a thoroughly bizarre situation, for unquestionably the thing already embodies Cartesian philosophy, and one could justifiably conclude that the Cartesians were a good three or four centuries late. II. And now to the issue at hand. Let’s talk about money. We have the seventeenth century to thank for the institution we flaunt today: namely, the central bank. Three markers are of significance here. With the founding of the Amsterdam Wisselbank (Bank of Exchange) in 1609, the time comes up with the prototype of a modern bank. The Swedish Riksbank, founded in 1656 and nationalized in 1668, constitutes the first national bank, also confronting us with the first assignat currency (albeit for a thoroughly base reason, the sheer lack of precious metals). This was followed at the end of the century, in 1696, by the much more famous Bank of England – the great model of the national central bank. It presents the image (at least seemingly) of how we understand a central bank today: a national institution which supplies money for a community with clearly marked borders: a money machine, a money engine. This means that the “Leviathan” is no longer a colossus on feet of clay, as Jean Bodin supposed, but a capital figure. All of a sudden it’s there, and with what verve! If you open an economics textbook you will find that the central bank is the a priori figure as such; it is taken for granted that there is an authority which regulates the money supply, and above all that there is no other supplier. Even Niklas Luhmann, who has a feel for economics, if not for history, speaks – in a psychological, un-Luhmann-like and thus noteworthy turn of phrase – of “the central bank” as “the ego of the system, so to speak.”3 The central bank, we may deduce, is for the state what the cogito is for modern philosophy and what the sun, as the center of gravity, is for Newtonian physics. By now we may have left the cogito and Newtonian physics in a shambles, but the central bank still stands pretty much unscathed. But why? Evidently because no one ever thinks of dismantling its privileges. But nothing would be easier . . . I, for example, instructed by my father that my genealogy goes back to a royal house, could start is3
Cf. Niklas Luhmann. Die Wirtschaft der Gesellschaft. Frankfurt a.M.: Suhrkamp, 1988. 147.
440
Martin Burckhardt
suing a Burckhardt sovereign, a perfectly reliable currency, stable in value . . . Who could stop me? This is exactly the case, or more precisely the dilemma, to which the central banks respond. By monopolizing the issuance of money, they present the monetary counterpart of what politics calls the monopoly on force. It was none other than Karl Marx who described money’s a priori structure by saying that “money wears a uniform and speaks different national languages.”4 Now, as we enter the precincts of the seventeenth century, this a priori can by no means be taken for granted. For money neither speaks the language of its issuers, nor does it wear a uniform (any more than anyone wore a uniform – Oliver Cromwell’s New Model Army was the first uniformed army). In the monetary sphere, chaos and anomie prevail. That – sheer desperation – is the reason for the founding of the Wisselbank. Just try to picture it: at the turn of the century in the Generalstaaten (that is, the area that is now Holland), there were not only two competing coins per province, but six municipal coins as well, for a total of 36 different coinage systems. At that time, shortly before the outbreak of the Thirty Years’ War, Germany entered the so-called “Kipper- und Wipperzeit” (the “Seesaw Era”), a period of devastating devaluation and currency debasement. The reason for this is easily determined: by that time the old coinage prerogative, that is, the royal privilege of the mint, had long since become a title which could be leased to second and third parties. In other words, what we now call junk bonds, derivates, futures, etc. was the norm. Any schmo could buy the coinage prerogative, cut off a slice for himself, sell it to another schmo, etc.5 The Amsterdam Wisselbank was such a great success because its account management made allowances for the monetary anomie and stopped ascribing any significance to any nominal value. Instead, sums were converted into an ideal currency that guaranteed the depositor that 4
5
“Money assumes a local and political character, it uses different national languages and wears different national uniforms . . .” Karl Marx. “A Contribution to the Critique of Political Economy. Part one.“ Collected Works. 50 vols. Moscow: Progress Publishers, 1975-2005. Vol. 29 (1987): Economic Works, 1857-61. 342. “The period from the end of the sixteenth century to the French Revolution is filled with currency upheavals and attempts to find ways out of them, which only had real, lasting success in England – where the problem was never as bad in the first place – elsewhere relief was only temporary. Of course, things were worst in Germany . . . The devastation of the Thirty Years’ War led to a huge coinage calamity in the very first quarter of the seventeenth century. That was the actual Kipper- und Wipperzeit, whose indescribable chaos is not easy to convey.” Joseph A. Schumpeter. Das Wesen des Geldes. Göttingen: Vandenhoeck & Ruprecht, 1970. 53.
The Illusion of Power: Central Bank Money
441
he would be given back the exact equivalent of the gold and silver he had deposited. So a virtue was made of necessity; in wise consideration of the fact that the country was swarming with counterfeiters, a shift was made to the metal value – but also to a highly abstract notation (back to the future). Let us jump to England for a moment. There the Stuarts, impoverished kings every one of them, had left the coin business in such disrepair that people who wanted to deposit money simply left it at the goldsmith’s. The goldsmiths then issued banknotes, or, actually, receipts for a certain type of coin. Like the Amsterdam Wisselbank’s ideal money, these receipts were soon circulating better than the coins themselves. Yet the goldsmiths were anything but altruistic. They used their skills to melt down the coins and reissue them with lower gold and silver contents – making the money supply rapidly double in size (and making prices reach hysterical heights). In 1696, moving toward a coin reform, it was determined that a pound, which was, after all, supposed to weigh one pound, averaged at only half, and in many cases only a quarter, of this ideal weight. Which illustrates the reality described by Gresham’s Law. Like tax evasion today, counterfeiting and coin debasement was a national sport. Counterfeiters were promised immunity from prosecution if they denounced two other counterfeiters, sons were explicitly encouraged to report their counterfeiting fathers – and those caught counterfeiting were executed. To make a long story short: if the monetary system had an outstanding feature, it was sheer, indescribable anomie. From this perspective, the institution of the central bank had a kind of divine impact. III. To give a better sense of the situation in the seventeenth century, let me take a quick look back at the history of money in Europe. The circulation of high-grade metal coins is a relatively recent phenomenon. In the mid-thirteenth century Genoa and Florence issued gold currencies, and Louis IX likewise hastened to bless his subjects with a gold and silver currency (which earned him sainthood and a louis d’or). It is not so farfetched to read this gilding process as a kind of monetary miracle of transubstantiation – or, more precisely, as proof of its rightness. Just as, in the miracle of transubstantiation, Christ enters the wafer, becomes the “true presence,” so do the ghost currencies of the Middle Ages (that is, the purely imaginary value descriptions customary in trade among
442
Martin Burckhardt
monasteries) suddenly become material: real presence.6 Miracle or no, there was a method to it – the very proto-capitalist method which Gothic Europe displayed so impressively during its cathedral boom. A superficial knowledge of the cathedral projects, with their high degree of mechanization and division of labor, the medieval textile industry, etc. makes it clear that the realization of the coin follows a certain intrinsic imperative – that a time which subscribes to the clockwork and the division of labor becomes increasingly dependent upon monetary mediation. In this respect, money’s realization has a certain logic. Once the money was issued, as economic historian Carlo Cipolla notes, it caused a chain reaction. Medieval society was seized by a kind of gold hunger, reflected not only in avaritia, but also in the Franciscan pageant of poverty. Even if money’s realization followed a certain logic, it must be said that medieval society was far from prepared for this logic. On the contrary: however perfect, money constituted a kind of foreign body – and here begins a strange, schizophrenic tale most easily told as a kind of Midas story. Money’s touch sets off one catastrophe after another. Relatively quickly Europe experiences a flourishing of usury – and the scholastic world can find no response to it, the Aristotelian ratio does not help in the slightest. What is worse: with respect to the hard core of materialism, the patron saint of logic and the natural sciences is a hopeless reactionary. In Aristotle’s expert opinion, money is “sterile, artificial wealth.” Hence it is inconceivable that it could produce children, i.e. interest, or more abstractly: added value, by itself (Greek has one and the same word for interest and child). But that is exactly what happens in usury – “pecuniam pecuniam parit” – and it was happening everywhere. The practice of usury touched on a kind of horror vacui. Twofold sacrilege, twofold sin: not only did it violate nature, above all – even graver – it violated the Christian conception of humanity. Charging interest, you see, was regarded as a form of “symbolic warfare”; the church father St. Ambrose teaches that interest payments are taken only from those against whom one should properly wage war.7 Moreover, this unnatural, un-Christian fertility conflicted with the internal logic of the medieval conception of money. Money was regarded not as a value in itself, but as a kind of calibrating instrument, a thermometer or a yard6
7
On the ghost currencies of the Middle Ages, cf. Carlo M. Cipolla. Money, Prices and Civilization in the Mediterranean World. Princeton: Princeton University Press, 1956. 42f. See the slender but outstanding work by Benjamin Nelson. The Idea of Usury. 2nd expanded edition. Chicago and London: University Press, 1969.
The Illusion of Power: Central Bank Money
443
stick that merely reflects the value of a thing rather than embodying it. For this reason scholars maintained that money is only loaned to the subjects, while the prince who issued the money (and whose portrait on the coin guarantees it) remains the money’s owner. It is evident, in view of the universal gold hunger, that this construction could not have much of a future. Now not only the usurers and their customers have succumbed to the gold hunger; the entire community is forced to watch its own evacuation to the extent that it follows the logic of money. First isolated, commercially active orders such as the Cistercians are released from tithing; then towns buy their freedom, finally purchasing the surrounding lands, parishes, etc. in order to achieve economic independence. Feudalism is becoming increasingly porous, eviscerated, and desubstantiated. All the same, the subjects demand that their sovereign defend the country, maintain an army, mount crusades, and so forth. So it was, barely twenty years after the introduction of the real presence, that the following situation arose. To finance his crusade, Edward I of England had borrowed money from the English Jews. The Jews were held in disfavor, and the more the usury dilemma became apparent, the more anti-Semitism increased. There were accusations of ritual murder, pogroms. The king (Edward, the Hammer of the Scots) took a stand, in 1275 making the Jews his subjects under the condition that they renounce usury and identify themselves by wearing the notorious yellow badge. However detestable this arrangement may seem from a modern perspective, it did provide the Jews with a legal status that had hitherto been denied them (the right to settle in cities and towns and to choose their profession). In return for royal protection, the Jews were obliged to pay taxes. This relationship between sovereign and subject was an innovation – in a formal sense, it anticipated the relationship that did not actually establish itself until the seventeenth century, after endless struggles. In fact, the statute of 1275 was not just an oppressive measure; rather, it protected the Jews from mob justice and created a legal relationship. The crucial goal was not achieved, however: usury proved impossible to forbid. Thus the agreement did not hold for long, and anti-Semitism only flared up all the more violently. In 1278 all Jewish households were searched, large numbers of Jews (but not Jews alone) were accused of coin debasement and hanged, while an additional 600 Jews were arrested and only released on ransom. In 1290, Edward, back from his crusade and deeply in debt to Venetian bankers, was forced to ask Parliament for permission to levy a special tax. The English barons, who had long regarded the Jews as a thorn in their eye, asked in return
444
Martin Burckhardt
for the expansion of parliamentary rights – and for the expulsion of the Jews. In contrast to the pogroms, this expulsion took a relatively civilized form; all the same, anti-Semitism had become institutionalized. That November all Jews were forced to leave England; anyone who remained was subject to the death penalty. Most went to France, but there they were banished once again by the French king, Philip le Bel. In England the Jews had been expelled and the Parliament had secured new rights. But usury remained; it had merely changed hands, was now a “Christian affair.” Henceforth there were increasing complaints about Christians disguising themselves as Jews to practice usury. In 1364 Edward III authorized the city of London to issue an Ordinatio contra Usurios, with a further edict following in 1390. To accommodate this foreign body, people now had to go about restructuring Heaven (the story Jacques le Goff tells in The Birth of Purgatory). In fact, in the fourteenth century the Midas touch actually caused the princes themselves to go into counterfeiting. Philip le Bel (the same king who banished the refugee Jews for the second time, and who, in the words of Marc Bloch, was the first to systematically turn the kingdom into coin) was a counterfeiter on a grand scale. In a sensational show trial, for the purpose of which the papal seat was moved to Avignon, he robbed the Templar Order, to which he was in debt, and had its Grand Master burnt at the stake. And he forced the pope to declare that the denial of the notion that usury was a deadly sin was itself a deadly sin.8 All in all, the fourteenth century can be called the age of the counterfeiter kings, with counterfeited coins, state-internal counterfeiter wars and the resulting rampant inflation. “Money rules the world” – however, money had not even begun to find a form that could guarantee its stability. It oscillated, metamorphosed, and deteriorated – and with it Christianity experienced a progressive devaluation of its values. Or, more precisely: slowly but surely it adjusted to the monetary realities. In 1487 England issued an edict against usury which, in fact, clandestinely sanctioned it by allowing for a poena conventionalis or usura punatoria. Finally Henry VIII legalized this state of affairs: he permitted interest payments, with a margin set at 10%. This legalization procedure was not without ulterior motives, however. By splitting with Rome, the king had taken possession of the church’s immense fortune,
8
Cf. Martin Burckhardt. Metamorphosen von Raum und Zeit. Eine Geschichte der Wahrnehmung. Frankfurt a.M.: Campus Verlag, 1994. 85ff.
The Illusion of Power: Central Bank Money
445
and now he wanted to invest it profitably (allowing the English feudal system to survive for another century and a half).9 If I fall back on the Midas figure to describe this process, it is because it cannot be comprehended as a rational sequence of events. From a historical point of view, Cipolla’s notion of the “chain reaction” is quite apt – with the caveat that in this case (involving a purely symbolic quantity) an image borrowed from natural processes is especially inappropriate. Above all, it evades the crucial problem: that this “chain reaction” is reflected in the reality, but not in the discourse. Indeed, the better part of scholastic erudition is devoted to the denial and suppression of this novel force – to not acknowledging the logic of money. So we have a discourse – a course – without a discourse, we have a discours de la methode which gains ground by itself, as it were, subcutaneously, clandestinely: a subterranean, headless system of desire. Countless works grapple with the “fair price” and condemn usury; the thinkers who confront reality as it is, not as it should be, can be counted on one hand, on two fingers.10 IV. The seventeenth century marks the point at which this unresolved conflict enters its last, apocalyptic stage. In this sense, the monetary confusion is merely a symptom of ideological anomie. This explains the ill humor which Thomas Hobbes, the great psychologist of civil war, displays in his Behemoth. When he denounces scholasticism as a kind of sinister sophistry whose only purpose is to “make money” out of the ignorance of the masses, he seems to be thinking mainly of his own contemporaries who intone the “political theology of the Middle Ages” while accepting payment in cash. This structural, schizoid bilingualism is what he is attacking – and if there is one point to his teachings, it is to melt down the scholastic contraband and institute the new, monetary, rational logic. Just one small example from the Leviathan: The Value, or WORTH of a man, is as of all other things, his Price; that is to say, so much as would be given for the use of his Power: and therefore is not absolute; but a thing dependant on the need and judgment of another. An able 9 10
Cf. R.T. Tawney. “The Rise of the Gentry.” History and Society. London, Henly and Boston: Routledge & Kegan Paul, 1978. 104. A rare exception, but all the more significant for that: Nicolas von Oresme. Traktat über Geldabwertungen. De Mutatione Monetarum: Tractatus. Epilogue by Martin Burckhardt. Berlin: Kadmos, 1999.
446
Martin Burckhardt
conductor of Souldiers, is of great Price in time of War present, or imminent; but in Peace not so. A learned and uncorrupt Judge, is much Worth in time of Peace; but not so much in War. And as in other things, so in men, not the seller, but the buyer determines the price. For let a man (as most men do) rate themselves as the highest Value they can; yet their true Value is no more than it is esteemed by others.11
If this quote sounds cynical, it is because the figure of thought is divested of all metaphysical trimmings, because we are presented not with references to humanitas, “dignity” and the like, but with the “utility value of the person” – to the extent that it represents exchange value (social esteem). Formulating this in Luhmann’s coinage – which is not usually my custom – one would say that the principle of “self-observation” has been switched here to “system observation.”12 That is indeed a fundamental switch – a radical shift in perspective. But is it correct for me to say that there is no metaphysical contraband involved here? Oddly enough, the text leaves one question open. By reducing everything to the price, Hobbes suggests that it must be backed by an absolute, permanent, and invariable agency: money. Not as a private engine of esteem, but as a social aggregate, as a universally binding reference value that permits mutual esteem. This universally binding nature, however, is exactly what cannot be assumed, on the contrary. Up to the end of the seventeenth century all rationality is mocked by a monetary system which anyone who knows his stuff can undercut as he pleases. In actual fact, money, or, more precisely, the inadequacy of the “universal equivalent,” poses the very metaphysical question that Hobbes tries to evade by bringing up the monopoly on force. How can a society of divergent and, when push comes to shove, even anomic individual interests manage to come to a consensus on one and the same Schein? How can a power succeed in producing the Schein? In essence, this question takes up a figure of thought originating back in the fourteenth century, that is, at the time of the most rampant counterfeiting, by the intellectual anomaly Nicole Oresme. Here the question goes like this: when the princes abuse their coinage prerogative by putting inferior money into circulation, they harm society. And one must ask: who does the money belong to? Oresme’s response is that the money belongs to society, that the prince only issues “the coin on its behalf and for its 11 12
Thomas Hobbes. Leviathan. Middlesex: Penguin Books, 1968. 151f. Although Luhmann always gives the question of money pride of place, as the crux of the social, so to speak, it is the great weakness of system theory. This is not because Luhmann fails to grasp the prominent significance of money, but because he assumes the institution of the central bank – as the “ego of the system” – a priori.
The Illusion of Power: Central Bank Money
447
good.” On this premise he conceives something like a central-perspectival society in which power has the task of “issuing the Schein” (albeit in a controlled fashion, that is, under supervision). Thus the question of money is used as a peg to anticipate an ultramodern social system. This is no mere theory – it is reflected by the Italian republics, Genoa or Florence of the early fifteenth century. Incidentally, it is no coincidence that the first picture entirely constructed around a mathematical central perspective is Massaccio’s “Tribute Money.” The background for all this is the reorganization of the Florentine land register. In an utterly modern fashion, all the citizens of Florence are entered into a kind of database, with clerks going from door to door and using a multiplechoice method to ask the sex (1 or 2), profession (denoted by a code), and assets of the interviewees and the number of mouths (bocche) they had to feed. Today you can go online and call up the financial circumstances of the Quattrocento in a database (the very database used by the city treasurers themselves). Thus money raises a question that pertains not to some purely procedural system, but to a specific form of rule. From this point of view, when the move was made to found the Bank of England in the 1690s it was only natural that voices would be raised claiming that such an institute would necessarily end up creating a republic. V. Before moving on to more theoretical reflections, let us examine several immediate considerations that played a crucial role in the Bank of England’s founding.13 Here I must mention two great minds who played a prominent role in its founding, though they are rarely cited in this connection: John Locke and Isaac Newton. Locke in particular (who wrote several theoretical discourses on money which are also worth reading from a semiotic point of view) was a key figure in the second act of the bank’s foundation, namely in the coinage reform. What was the issue there? The crucial desideratum was to equip the state with a reliable, calculable coinage system. This was prompted not only by the experience of Charles I, the unfortunate debtor king who died by the axe as king in the king’s name, but also by all the financial catastrophes the kingdom was going through. In the fourteenth century it was a mat13
Cf. Andreas Michael Andréadès. History of the Bank of England 1640-1903. London: P.S. King, 1966.
448
Martin Burckhardt
ter of protecting the population from their own prince; by the seventeenth century the battle zone had expanded, the desideratum was to protect the state from being plundered by its own inhabitants. The goldsmiths, for instance, had not hesitated to charge as much as 25% interest when they lent money to Protector Cromwell and later to the king – which in 1672 led to national bankruptcy. Thus the first goal was to fix the interest rate, and the second – now that paper was preferred to the corrupted coins – was to begin issuing banknotes as well. Nominally speaking, the bank was a private institution. It was authorized to trade not in goods, but in bills of exchange, gold and silver bullion; in addition, it was allowed to sell all the goods that had been deposited as security but not redeemed (i.e. property, ship cargoes, etc.). What it had in common with the Swedish Riksbank, and what distinguished it from its historical predecessors in Venice, Genoa, and Amsterdam, was the fact that the Bank of England was able to loan money. This was a crucial new phenomenon: while earlier banks had acted as a kind of container, holding money in safekeeping (and receiving payment for that) and trying to increase their capital by equipping various ventures, the Bank of England was authorized to loan money (i.e., in fact to create money). The first loan was to the English state, which agreed to pay off the debt at a certain fixed interest rate. All the same, the Bank of England fell short of previous achievements. In Venice, for example, the Rialto Bank’s banknotes had been made legal tender, while a similar procedure was adopted in Genoa in 1675. In fact, this inconsistency shows that the status of a central-perspectival institution of this kind had not yet been clearly defined. Accordingly, a so-called “Land Bank” was founded not long after on the principle that real estate also represented a kind of bound, immovable capital. Yet this institution had no luck, failing miserably – in large part because the advantage of coin, as opposed to territory, lies in its mobility. This collapse strengthened the Bank of England, which was now promised a twenty-year bank monopoly (which, after positive experiences, was soon renewed and extended). On the whole it can be said that the institution’s structure was only developed through a process of trial and error. Incidentally, this corresponds fairly exactly to the revolution that took place in our lifetimes: namely, the introduction of free floating. Then too – in the early seventies – it was far from certain whether and how such a mechanism would function: initially it was believed that exchange rates could be agreed upon and maintained (an illusion which the international financial markets rapidly destroyed). The founding of the Bank of England was a similar case: the instruments which we ultimately regard as constitutive
The Illusion of Power: Central Bank Money
449
of a central bank only developed over the course of time. Nonetheless – and despite adverse political conditions – the foundation of the bank was a great success. Quite clearly, the populace was ready for the Schein. This Schein, however, is separate from the position of the king. If one compares this with the ordeals that William III had had to go through just a few years before, the caesura – indeed, a virtual revolution – becomes apparent. Despite the threat posed by Louis XIV’s troops, the king was no longer creditworthy; no one, neither the goldsmiths nor the merchants, was now willing to lend the king money to defend the country militarily. The king’s representatives had to go from house to house and shop to shop like beggars to scrape together the money for the country’s defense. VI. Given the central bank’s past history, one cannot say that anything ground breakingly new, in a positive sense, was happening here. Quite some time ago the Venetian Banco del Rialto and the Genoese Banco di Giro had already anticipated what was realized in the Amsterdam Wisselbank, the Swedish Riksbank, and the Bank of England. As an intellectual construct, such an institution would have already been conceivable in the fourteenth century (and at the time Oresme’s monetary theory already called for such an institution). In this respect it strikes me as highly dubious to designate the seventeenth century – in terms of a positive innovation – as the beginning point of our social system and way of thinking. The mechanical clock, the monetary change, the printing press, central perspective – all these instruments had long since come into being. Thus one could ask, conversely, which instrument, if any, we can attribute to this century’s ingenuity. (While Hans Blumenberg selects the telescope and ascribes such paramount philosophical significance to the telescopic gaze, I do not feel that this was the most felicitous choice, given that the innovations mentioned above are of much greater complexity and above all more powerful in effect than the telescopic gaze.) Capitalism’s discours de la méthode is an apocrypha, a dubious text – but, for that very reason: the collective phantasm. It seems to me – and this brings me back to the point which I mentioned at the beginning and which goes far beyond the founding of a political institution – it seems to me that the special achievement of the seventeenth century lies not so much in the creation of new instruments as in the systemic completion and implementation of the resulting fig-
450
Martin Burckhardt
ures of thought. In the case of the emerging central bank, it is the collective willingness to trust the Schein – without there being a stigma attached to money or monetary production, as was still the case in the fourteenth century. The credit system (or, to formulate it more generally, the Schein) has come into being. If there is one thing we must give that time credit for, it is that it exorcised the Middle Ages in an extremely violent and occasionally catastrophic manner. While the doctrine of transubstantiation embroiled medieval thought in the conflict of political theology, meaning that a genuinely political discourse was continually being negotiated in a metaphysical form, seventeenth century thought arrives at a monopoly of the central perspective and thus at a completion. Now one speaks not of epiphany, but of the machine; one speaks (like Hobbes) of God as a “legal entity” which can be represented only by the state, the Mortall God. The strategy is obvious: the position of God, the discourse of theology, is hedged in, camouflaged. In contrast to medieval theology, the king no longer has “two bodies” (a transcendental and an earthly), he has only one (though it is coded in two ways: as the “body of the office” on the one hand, and as a private person on the other). The “Leviathan,” the central bank, the standing army, universal taxation – all this can be regarded as the doctrine of transubstantiation taken to its logical conclusion. The medieval belief system has turned into a secular system of credit fallen into the hands of a monopolizing nation-state. And so the reason for this exorcism is threefold: first, it drives out the rival belief system (preventing a relapse into a theologically-based order); second, the position of God is centralized and pressed into service; and third, the genealogy is garbled, creating the illusion of selfproduction, of autopoiesis. We – and this is the weak point of someone like Edmund Husserl, pondering the crisis of European sciences – perceive only this third level. Thus we speak of innovation and renewal when one could say with equal justification: “There was more ending than every before!” Medieval thought had come to an end, and suddenly things – or more precisely, suddenly the long-existent ratio – could explode, and at the same time long-outlived traditions could be jettisoned. That is what I mean by an “a priori completion.” Rather than speaking of a “revolution,” I prefer to speak of the notorious match that sparks a revolution.
The Illusion of Power: Central Bank Money
451
VII. If one approaches history from the position of the historian, that is, the “prophet in reverse,” the institution of the central bank is probably one of the most important facts of the seventeenth century. At the same time, it would be shortsighted to detach this institution from the power aggregate that formed in this period – the nexus of money, power, and centralized force that I will refer to as the “state machine.” In a certain sense, at least in retrospect, this state machine appears as a kind of organization based on natural right. When Marx writes that money “uses different national languages and wears different national uniforms,” he is referring to this aggregate (and at the same time immortalizing it). From this point of view the state machine no longer appears as a historical and thus vulnerable construct, but as a kind of super-historical monolith, an a priori battery. Well and good, you might say: we know Marx didn’t understand the first thing about capital, but what do we have political economy for? Let me give an example. In 1977, that is, not long after the oil crisis, and above all at a time in which the gold-backed currency had given way to a free-floating one, the economist Friedrich von Hayek, already a venerable gentleman of seventy, wrote a little book called The Denationalization of Money.14 In this work Hayek expresses his surprise at the fact that the “existing literature . . . [offers] no answer to the question as to why a government monopoly on the money supply is seen everywhere as indispensable”15 – or in other words: why money, specifically central bank money, is assumed a priori. In fact, political economy always assumes as a given what it purports to eliminate. Here we are confronted with the peculiar ambiguity I had in mind when I mentioned the Schein. The state machinery is a Schein-producing apparatus in a twofold sense. Not only is this machine granted the right to issue money, an additional level of illusion is touched upon: that no one recalls that it was not this way from the beginning. An Adornite, schooled in the history of catastrophes, would probably speak of a “general context of delusion.” Before taking refuge in such a desperate formula – and one so highly saturated with ideology – I would like to approach the problem more rationally. The first key point is the startling phenomenon that the banknote entails a different form of the Schein. What can we call this? 14 15
Friedrich von Hayek. Die Entnationalisierung des Geldes. Tübingen: Mohn, 1977. Ibid. 6.
452
Martin Burckhardt
Illusion? Or money illusion, as the economists suggest. The appeal to the illusionary is anything but helpful, for collective illusion represents an emergency, not some kind of an amusement park. The Schein becomes real – the same way in which the Lateran Council asserts Christ’s real presence. But how does this happen? Let me take a stab at describing the mechanism. When, seen a posteriori, the central bank takes a priori form, we have to say that a historical completion seems to be taking place. The question answered by the central bank no longer arises. In this respect, the other side of the Schein is a historical obfuscation, a feat of suppression all the more remarkable for the fact that it involves not just a given historical collective, but an entire epoch. VIII. In conclusion, let me examine the other side of the Schein. We know that sociology, following Pierre Bourdieu and his “fine distinctions,” speaks of profits of distinction, that is, it operates with a universal currency of prestige. If we trace back the coinage of this thought, we come to Norbert Elias and his study of court society. The title is somewhat misleading, as Elias is concerned not with “court society” in general, but only with the court of Versailles and court society of the seventeenth century – but so be it. What Elias elaborates here is the “illusory currency” of prestige, the peculiar apparatus referred to even at the time as le mecanique, the mechanism (Elias speaks of a perpetuum mobile16, a prestige fetish, etc. – thus taking refuge himself in peculiarly irrational figures of thought). Yet if one regards Versailles as a “medial machine,” one sees that this currency of prestige is anything but arbitrary; rather, this form of symbolic capital is analogous to what occurred when money was centralized in the central bank. In this respect this is a “machine of central perspective,” a centralization of power in the actor playing the king, a centralization of force in the standing, uniformed army, a centralization of money in the bank note. You see the same structure everywhere. Versailles is interesting (and definitely more amusing) to the extent that it demonstrates the genealogy of the Schein. What happened in Versailles was the landscaping of the central perspective. The abstract laws of image production were naturalized, creating the illusory “second nature” which replaces the hair with the wig, the complexion with 16
Cf. Norbert Elias. The Court Society. Oxford: Basil Blackwell, 1983. 86f.
The Illusion of Power: Central Bank Money
453
powder, perspiration with perfume. The fact that Louis XIV (a thoroughly mediocre man) managed to lure the landed gentry to the court of Versailles not by using brute force, but through a display of power at the court, proves that the mechanism of representation itself must be read as a structure of desire, not of force. In this respect it cannot be imposed upon anyone, but only put into service. In fact this structure of desire (like the desire for interest-producing money) has a tradition – anchored in people’s minds like a kind of private religion – going back not merely to the seventeenth century, but much, much earlier. The French king’s stroke of genius was to provide this structure of desire with the matching ambience, the architecture and the environment that made it possible to centralize desire, to politicize it, to transform it into calculation. Aesthetics is pressed into service to create the political mechanism – the symbolic monetary system that regulates access to power. In this sense, the currency of prestige should not be seen as an arbitrary, symbolic currency; it is dependent upon a specific grammar (an argument which ought to be raised against Bourdieu). In this respect, when I speak of the instrument of the Schein, I mean it very much in a strict sense, not a loose one. I am saying that the logic of the central perspective, the code of representation is hidden behind it. In the seventeenth century the instrument of the Schein acquires a new significance, however. Rather than standing in front of the pictures, one moves within the “picture,” one takes a standpoint, one is, as it were, enveloped by Schein. This is made possible because in the seventeenth century the Schein gains the monopoly – and the nationalization, the collectivization of this logic begins. From now on the Schein can be counted on (as noted above, Sweden circulated assignats, or bank deposit money). Being in the picture means two things: it means that one can count on the laws of the picture while at the same time forgetting that one is subject to “the logic of the picture.” This nationalization institutes a structure of desire as a universally binding order – and subjects violations of the Schein to draconian punishments. The sovereign is the one who is able to supply society with the Schein it needs – in two senses of the word, as aesthetic appearance and as a banknote. In this respect sovereignty is no abstract magnitude; it is under constant trusteeship. Its yardstick is the subjects’ willingness to take the illusion at face value (and the great wisdom of Louis XIV consisted in legitimating power by stage-managing the central-perspectival language of hegemony – turning the trompe l’œil of representation into the state religion). In fact, these two aspects of the Schein overlap – it is fundamentally impossible to separate the two sides of the coin. Just as
454
Martin Burckhardt
Christ’s double nature has both a fleshly and a pneumatic-divine side, the Schein has a profane and a “divine” side: image and number. IX. Now, I fear I will exceed my allotted time if I succumb to temptation and continue pursuing this line of thought. I would like to add one point, however, in part because the above-cited sociologists have provided us here with circularly reasoned, tautological machines (be it the prestige machine, the distinction machine, or the social machine) – machines which are capable of all sorts of things, but which founder hopelessly upon money’s circular form. Let us take one exhausted player, let us take John Searle, who bit off more of the money question than he could chew in his Construction of Social Reality. Searle is user-friendly enough to formalize the problem for us, coming up with statements such as Z counts in System C as a representation of Y(X) – he calls that the “representation formula” Z counts in System C as a representation of Y(X) where there is no X – he calls that the “fiction formula” We accept (S has power (S does A)). – “formula for conventional power” We accept (S is trustworthy (S does A)). – “the informal trust formula”
Perfectly nice formulas, but completely inadequate, for they do not explain how representation or fiction, power or trustworthiness comes about. They describe and formalize what already exists within the system (and in this respect they are mere tautologies). If you take a close look, you will see that what Searle breaks down into separate formulas is actually an entire complex: If a system has been instituted in which something represents something else, it is because it has been issued by a certain trustworthy power. Power, money, trustworthiness – all this joins to form an effective illusion, this miracle of transubstantiation, European-style: that the Schein becomes reality. That can be described, but not (in the sense of a chess game) formalized. Otherwise I could whip out my royal coinage prerogative (Charlemagne, as you may recall) and issue my Burckhardt – but, as you can well imagine, this would be an undertaking with little prospect of success. It might prove that the money token is an “arbitrary sign,” but I would still be unable to give my money token validity. The crucial question about the Schein is: Why does this mechanism, the social game of money, function so smoothly? How has money, “this nothing kept in short supply,” become an all-purpose glue holding together an
The Illusion of Power: Central Bank Money
455
infinitely complex social system? This brings us to the question of Schein not as rationality, but as a phantasm (in other words, to the other, irrational side of the ratio). I said above that the machine of the state takes God’s position and immediately (like the dwarf in Wolfgang von Kempelen’s chess apparatus) becomes invisible. That, I fear, is characteristic of what happens with the “a priori completion”: a “metaphysical question” is encapsulated. Where then, you may ask, is the dwarf hidden, the little hunchback of theology? Where indeed. In the center, of course. I spoke of “representation” – and Searle also speaks of representation, in the highly restricted sense customary only since Hobbes: in the sense of mere substitution, metonymy. What kind of a dwarf, do you think, could be hiding in this term? Well, in the Middle Ages, for Thomas Aquinas, repraesentatio meant the “register of the saved souls in the book of life in Heaven,” or, translated into contemporary lingo, the “celestial database.” To use a term such as representation without any awareness of its dark, opaque side is to be taken in by the phantasm of reason, to be “blinded by reason,” as it were. But this very blindness is the point of money, in this blindness the Schein circulates, because we believe in it – but what do we actually believe in when we believe in it? In reason? In the state? In force? I am the very last person you should be asking. I picture my father – my family tree – with Charlemagne at the very top – and think to myself: that that is a solution I could live with. Trust yourself – trust your fake. Thank you for your interest. Translation: Isabel Cole
456
Martin Burckhardt
WORKS CITED Andréadès, Andreas Michael. History of the Bank of England 1640-1903. London: P.S. King, 1966. Burckhardt, Martin. Metamorphosen von Raum und Zeit. Eine Geschichte der Wahrnehmung. Frankfurt a.M.: Campus Verlag, 1994. Cipolla, Carlo M. Money, Prices, and Civilization in the Mediterranean World. Princeton: Princeton University Press, 1956. Elias, Norbert. The Court Society. Oxford: Basil Blackwell, 1983. Hayek, Friedrich von. Die Entnationalisierung des Geldes. Tübingen: Mohn, 1977. Hobbes, Thomas. Leviathan. Middlesex: Penguin Books, 1968. Luhmann, Niklas. Die Wirtschaft der Gesellschaft. Frankfurt a.M.: Suhrkamp, 1988. Marx, Karl. “A Contribution to the Critique of Political Economy. Part One.” Collected Works. 50 vols. Moscow: Progress Publishers, 1975-2005. Vol. 29 (1987): Economic Works, 1857-61. 257-420. Nelson, Benjamin. The Idea of Usury. 2nd expanded edition. Chicago and London: University Press, 1969. Nicolas von Oresme. Traktat über Geldabwertungen. De Mutatione Monetarum: Tractatus. Epilogue by Martin Burckhardt. Berlin: Kadmos, 1999. Schumpeter, Joseph A. Das Wesen des Geldes. Göttingen: Vandenhoeck & Ruprecht, 1970. Searle, John R. The Construction of Social Reality. New York: Free Press, 1995. Tawney, R.T. “The Rise of the Gentry.” History and Society. London, Henly, and Boston: Routledge & Kegan Paul, 1978. 85-128.
SYBILLE KRÄMER
The Productivity of Blanks: On the Mathematical Zero and the Vanishing Point in Central Perspective. Remarks on the Convergences between Science and Art in the Early Modern Period
1. Instruments as Symbol-Technical Hybrids We usually associate ‘instruments’ with man-made products such as tools, appliances, machines, or devices. Ranging from the screwdriver to the piano, up to the telescope and computer, we create objects which are then put to use to achieve particular goals. Speaking concretely, such an instrumentality makes up the core of our technical relation to the world. This rests on a further conviction: if the technical is fundamentally related to the instrumental, then the symbolic, that is, the use of signs, is not instrumental but should be understood as oriented towards understanding and interpretation. The materiality and sensual thing-ness, which is also part of the sign, is not then articulated in the termini of opaque objecthood, but on the foil of a transparent, transitory mediality. Here our understanding of instrumentality seems to establish a dividing line: the technical relates to the symbolic as production to representation (Herstellen to Darstellen), construction to interpretation, and as instrument to medium. Technical and symbolic procedures embody, therefore, two distinct domains of human poiesis, each with its own procedural and developmental logics. Nevertheless, there are quite a few artefacts that block this schema of a disjunctive sorting into the technical or semiotic. Phonetic writing is a mechanism for setting down oral language – but is writing a technical or symbolic system? The decimal position system together with calculative algorithms allows a human calculator to ‘mechanically’ solve all the tasks of elementary arithmetic – so is the decimal system a calculating instrument or a numerical language? Is software, for example a
458
Sybille Krämer
computer program that transforms a physical device into an Internet portal, now a machine or a text? The list of questions could easily go on. Particularly in the field of cultural techniques, we stumble on such mixed forms that we want to call ‘symbolic-technical hybrids.’ Here, ‘hybrid’ relates to a combination of attributes originally belonging to distinct classes or fields of objects, which, however, cannot be distinguished from their ‘mixed being’; rather, they exist simultaneously next to each other, and in this way remain sustained in their heterogeneity. The hybrid does not follow the logic of ‘either-or’1 but of ‘both-and.’ To come back to our examples, phonetic writing, the decimal system and its calculation rules, a computer program, are all in fact both: technology and language, instrument and medium. Seen from the viewpoint of such hybrids, ‘technology’ and ‘symbol’ become purifying stylizations, pole and borderline-case, of a scale (still) only grasped as a concept-grid, whose spectrum delivers what in the world we only encounter, in reality, as mixed relationships. Our supposition, and also our hypothesis, is that lasting shifts or even ‘leaps’ of a cultural dynamic touch on a – in cultural-technical terms – definable ‘instrumentality’ of just such symbolic-technical mixed forms. This hypothesis concerns the foil on which convergences between instrumental innovations in science and art at the beginning of the modern period can emerge. 2. First an Anecdote Let’s start with a short story that Giorgio Vasari tells about Giotto.2 The pope wants evidence of the art of Giotto di Bondone, a contemporary of Dante. When a courtier asks Giotto for a sample of his skill, Giotto takes a brush dipped in paint, and using his arm pressed to his side as a compass, paints a circle whose perfection causes much amazement. He passes the drawing on to the courtier, who is unsettled by the meagerness of the sample. The pope however understands and recognizes Giotto’s genius. 1
2
On the hybrid in Bakhtin’s literary theory, cf. Irmela Schneider. “Von der Vielsprachigkeit zur ‘Kunst der Hybridation.’ Diskurse des Hybriden.” Hybridkultur. Medien, Netze, Künste. Ed. idem and Christian W. Thomsen. Cologne: Wienand, 1997. 20. Giorgio Vasari. The Lives of the Artists. Trans. Julia Conway Bondanella and Peter Bondanella. Oxford: Oxford University Press, 1998. 15-36. Concerning this anecdote, cf. Reinhard Steiner. “Malerei als Spekulation.” Digitaler Schein. Ästhetik der elektronischen Medien. Ed. Florian Rötzer. Frankfurt a.M.: Suhrkamp, 1991. 435f.
The Productivity of Blanks
459
What is revealing in this anecdote, which crystallizes a rupture in the fine arts into a legend? For a start, in relation to painting, drawing is revaluated. Moreover, the skill of the artist is demonstrated by the manner in which he is able to use his body as an instrument. Using his arm to imitate a compass, Giotto links ars and techné, painting and construction in the creative act. But what is particularly significant is that the painted image no longer follows the model of nature, i.e. a depiction, but sketches a mathematical object. Circles, in a literal sense, do not appear in nature – they are mathematical constructs. Plato was right when he insisted that the circle is not an element of the real world, but belongs to the world of the intellect,3 since the circle drawn as a geometric figure is the sensible embodiment of a concept, a formula, and consequently a theoretical and abstract object. Therefore the painter’s skill was directed towards a subject whose ‘nature’ it is to embody an object which can be constructed using calculations. 3. Calculability as a Symptom of Modern Art and Science, the Visualization of the Invisible, and the ‘Discovery of the Subject’ as the Origin of Numerical and Visual Space When we speak of a convergence in the development of art and science in the early modern period, it is calculability that declines the mutual measure of both. The invention and dissemination of linear perspective on the one hand, and the language of formulas on the other, can both be understood as a symbolic form that seals a sublime connection.4 Central perspective rationalizes our visuality; calculations rationalize our language. The construction of central-perspectival space sets out to objectify the phenomenon of seeing. The laws of perception are transformed into methodical and calculable rules of symbolic representation. The construction of calculations attempts to make operative the phenomenon of speaking, with the invention of a formal writing – this avoids vagaries in the use of natural languages and enables those who control the artificial alphabet of a calculation and its rules to identify errors in the representation as errors of calculation, and therefore to avoid them. 3 4
Cf. Sybille Krämer. Berechenbare Vernunft. Kalkül und Rationalismus im 17. Jahrhundert. Berlin and New York: de Gruyter, 1991. 53ff. Cf. Sybille Krämer. “Die Rationalisierung der Visualität und die Visualisierung der Ratio. Zentralperspektive und Kalkül als Kulturtechniken des ‘geistigen Auges.’” Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst. Ed. Helmar Schramm et al. Berlin: Dahlem University Press, 2003. 50-67.
460
Sybille Krämer
The ‘perception of something’ is made operative with the help of central perspective, just as ‘speaking about something’ is with the help of calculations. Consequently, what in artistic and scientific representation is made visible becomes in turn a mathematizable construction, and therefore a theoretical entity. As soon as we describe an object as a ‘theoretical entity,’ it belongs – should we use a traditional Greek distinction – to the realm of ‘logos’ and not to that of ‘aisthesis.’ The ‘logoi’ themselves are ‘anaisthetic,’ that is, invisible. Only accessible to thought, they belong – at least in the traditional sense – to an ideal domain, which is outside space and time. If however central-perspectival and calculative representations make theoretical entities physically present, then both cultural techniques can also be viewed as techniques for visualizing the invisible. And that is the point towards which we want to direct our attention here. The arts are normally associated with the sensual and the concrete, while the sciences are associated with what can be formalized, quantified, and with abstraction. Our assumption, however, of a convergence between art and science is rooted in the conviction that calculations, in which the cognitive entities appear in the two-dimensionality of formulas, also makes use of the potential of a visualizing of the non-concrete and invisible. The interplay of invisibility and visualization, belonging to the cultural techniques of central perspective and calculation, now corresponds to an interrelationship between the symbolic and the technical, since it is the (symbolic) act of the representation of an invisibility that can also be reconstructed as a (technical) act of the production of the represented. Now there is a phenomenon which is especially suited to demonstrate this connection of the sensualization of a non-sensual thing, which is only constituted at all as a result of this visualization, and that is the use of the figure ‘0’ in the context of arithmetical calculations, in whose wake the cipher zero first comes to ‘life’ as a mathematical number; and it is also the use of the vanishing point, the ‘visual zero’5 in linear perspective, in relation to which the observer’s point of view is promoted to the constructing principle of the picture. What unites both of these forms – at least this is our supposition – is that the world of numbers as well as the world of images are structured such that, from this point of view, an epistemological subject can be made imaginable as the origin of numerical space as well as visual space. Before we turn to the reasons for this supposition, a brief interjection seems in order.
5
Brian Rotman. Signifying Nothing. Basingstoke: McMillan, 1987. 19.
The Productivity of Blanks
461
4. Interjection: Is the New Scientific Research a Contribution to the Performative ‘avant la lettre’? The idea that, with signs, something can not only be represented but also produced is a core idea of the concept of performativity. In his pioneering work How To Do Things With Words, John L. Austin revealed that there are verbal expressions which concurrently also carry out what they say:6 a ship’s blessing, a judge’s ruling, the ‘I do’ of a marriage ceremony, testaments, but also promises, requests, orders. With the discovery of the performativity of linguistic expressions, Austin demonstrated a quasi-instrumental dimension in our speech: speech-acts create – at least under certain conditions – social facts. With their world-changing and world-forming power, verbal operations hardly come second to technical operations. This ‘instrumental revaluation’ of our speech by speech-act theory subsequently effected linguistics, philosophy, art, and cultural theory. On the threshold of the twenty-first century, a ‘performative turn’ is looming.7 Yet the theory and history of science (still) hardly refer to the performative, although the basic role of language as a means of presentation in science has become a commonplace. The fact that the linguistic aspect has fallen into the shadows has an understandable cause, since one of the most far-reaching changes of direction in scientific research intended that the notion of ‘science’ should not be reduced to the creation and verification of the theoretical forms of knowledge, that is, the propositional, and therefore to what is verbally explicable. As a result, non-verbal phenomena shifted into the focus of scientific research, such as the semiotically or technically supported practices in laboratories, the interaction of visuality and knowledge, image and text, the knowledge implicit in the use of scientific objects, the hybridization of things, symbols, and technologies in the day-to-day practice of the scientists. As much as we can discover in these approaches a reflection of the performative in science ‘avant la lettre,’ its determining feature remains a turning to the non-verbal aspect of the scientific process. It is therefore time to test what insights are disclosed as soon as the idea of linguistic performativity is removed from its original domain, in colloquial language, and made fruitful for the observation of the role of scientific languages. The operative writing of calculations, i.e. formula6 7
John L. Austin. How to Do Things with Words. Cambridge: Harvard University Press, 2005. Uwe Wirth, ed. Performanz. Zwischen Sprachphilosophie und Kulturwissenschaften. Frankfurt a.M.: Suhrkamp, 2002.
462
Sybille Krämer
language, is one of the most lasting forms of linguistic expression in the sciences. Where does it lead, if we now ask about the performativity of calculation, i.e. formal languages? Here we have come back to our theme. 5. On the Performativity of Calculation: When the Description Becomes the Execution of the Described From a performative perspective, we need to clarify how to understand the fact that calculated expressions represent something, and also give rise to the represented. To begin with, such a thesis finds itself in opposition to a readiness, which is common in the cultural sciences and inspired by debates in scientific research, to interpret the empirically operating and formalizing sciences not only as realistic, but also as instrumental: ‘instrumental’ in the sense that the statements of these sciences should not be interpreted as the depiction of an existing reality, but rather as the result of the interaction of people with the object-area to be investigated, so that it is legitimate to say that the things these sciences deal with are constructs arising as a result of scientific practices. Realism and instrumentalism contradict each other in relation to the question of whether theoretical entities actually exist (realism) or do not exist (instrumentalism). If we now approach this question in the context of a performative perspective, then we hope to be able to be in agreement with realism, in that the objects denoted by correct calculation-expressions actually exist, but that these objects are to be interpreted as the referents of calculations; these then – as instrumentalism supposes – do not exist independently of the symbolic mathematical practices. Here a revealing analogy for the question of the performative arises: just as performativity can create ‘social facts’ from (colloquial) speech-acts, that is, ‘objects’ whose existence is rooted in their social acknowledgement,8 the performativity of calculation creates epistemological facts, whose existence equally depends on their being acknowledged. Let us therefore turn to the symbolic act of formalizing. So-called formal languages are in reality writings: graphic systems sui generis, which then, in accordance with various colloquial languages, can also be uttered. The writing-character of ‘formal language’ is not a marginal characteristic but an essential one. Normally ‘writing’ is associated 8
Cf. John Searle. The Construction of Social Reality. New York: Free Press, 1995.
The Productivity of Blanks
463
with written-down oral language. Here however, a more broadly applied notion of writing is called for which does not follow a phonetic writing-conception but views writing as a kind of hybrid-construction between language and image. In this way a phenomenon arises which can be too readily overlooked and forgotten in an observation reduced to the discursiveness of writing – ‘Schriftbildlichkeit’ (a hybrid of writing and image).9 Unlike oral speech, which is dependent on the sequential nature of time, writing works with spatial configurations that make use of the two dimensionality of the surface. This spatiality opens up a kind of ‘in-between-spatiality.’ Schriftbilder are discrete, they work with gaps and empty zones, because of which letters first of all become unambigous, that is, disjunctively individualizable; no third sign can be placed between two signs. This exclusion of continuum on the part of the medium is the condition for the fact that we consider the represented in the medium as a system put together from identifiable elements. In the course of such a syntactically organized systematicity of the medium of representation, the represented takes on nolens volens systematic features itself. The visualization in the structure-image of writing subjugates what is visualized to the grid of a systematic construction principle. And we will see later, taking the numerical calculations of the decimal position system as an example, how important this transformation of numbers into a number-system, by means of notation in the form of a calculation, is; particularly when it is concerned with clarifying the meaning of zero. If we also want to understand the calculated representation as the production of the represented, then at best we have so far stumbled on one transformation. A transformation that consists in the fact that the visualizing representation of contents in the disjunctive Schriftbild (writing-image) of calculation reveals this content precisely in its systematicity. If we speak of ‘production’ here, then this only relates, more precisely, to a commutation. Hence we must take a further step; and only with this step do we stumble on a phenomenon that is revealed, exclusively, in the writing of calculations. Our previous considerations relate to the implicit structure-‘producing’ iconicity in calculation; now we want to turn to the operativity that is part of calculation. A calculus hides a double function: it is both language and technology, medium and instrument. In as 9
Cf. Sybille Krämer. “‘Schriftbildlichkeit’ oder: Über eine (fast) vergessene Dimension der Schrift.” Bild, Schrift, Zahl. Ed. idem and Horst Bredekamp. Munich: Fink, 2003. 157-76.
464
Sybille Krämer
much as decimal figures represent numbers, these figures are a numerical language; in as much as we calculate using writing, with the help of numeral configurations, decimal calculation creates an arithmetical problem-solving machine, and is therefore a calculating instrument. Unlike phonetic writing, which is already familiar to us, a Janus-headedness comes into play in the operative writing of calculation: the word ‘worm’ describes a worm, but it is not a worm; the expression ‘4 – 4 = 0’ not only describes a calculative operation, it is a calculative operation. Description and execution come together in the operative writing of calculation. We can also explain this coincidence in a different way. Before the introduction of decimal calculation, numbers were written in Roman numerals. However – since Roman numerals do not create a calculus but are only a numerical language – they needed to be calculated with a concrete reckoning instrument: the abacus. The medium of numerical presentation and the instrument of numerical calculation were realized in different ways. The ingeniousness of the decimal place-value system – making written calculation possible and the abacus redundant – consists of being able to realize the representing of numbers, and operating with numbers, within one and the same system. And this ‘double occupancy’ of a symbolic praxis – as description and calculation – is a characteristic of a calculus alone. Only the (correctly) calculated expression also carries out what it describes. And this marks the core of its performative dimension, belonging to logic, mathematics, the sciences, to computer programs, or to everyday calculations. In relation to the calculus, we have so far distinguished two aspects of its performativity. In the course of the medialization by the structureimage of the calculized writing, its referent takes on the status of a systematic entity, which is also a theoretical entity. In the course of the instrumentalization of the writing-medium as calculating-procedure, presentation and execution coincide. 6. The Example of Zero Using the example of zero, we want to show how, by the symbolic practice of written calculations in the decimal position system, a new, an epistemological object, is created. The hub of this consideration is a distinction. Normally we assume that the figure ‘0’ ‘means’ the number zero. Actually however the figure ‘0’ was invented centuries before the number zero was recognized as a mathematical object.10 The difference 10
Cf. “The invention of zero preceded its discovery by centuries.” Constance Reid.
The Productivity of Blanks
465
between figure and number is not simply the difference between a physically visible mark and its ‘invisible’ referent, but rather the different function which the figure zero adopts in relation to arithmetical signs. Intuitively we can clarify this by two expressions: In the equation ’10 – 1 = 9’ the sign ‘0’ serves within a numeral-expression to mark an empty place which shows the absence of one of the other figures (1, . . . ,9) in this position; here the ‘0’ functions as a ‘gap-sign.’ In the expression ‘0 – 1 = -1,’ on the other hand, we treat the figure ‘0’ as an independent number; it functions here as a number-sign. In the first instance ‘0’ symbolizes the absence of other signs, and in the second, the presence of a number.11 What we want to show now is that the creation and consolidation of zero as a mathematical entity is not already carried out with the introduction of the gap-sign, it is first ushered in by calculations with the figure ‘0.’ It is therefore a matter of the ‘birth of the number zero from the spirit of the cultural technique of the written calculation.’ The Greeks and the Romans did not know a zero. We ‘have’ numbers only in accordance with a medium, that is, a numerical language which makes the (invisible) numbers appear in a more material form, and then also makes them usable. However, the numerical languages common in Greek and Roman cultural circles – Greek letters and Roman numerals – treat numbers in the sense of a group of countable quantities, and this in such a way as to make calculating with these written numerical signs unsuitable. Such an identification of numbers with ‘quantity’ is then enforced by the praxis of using a physical calculating instrument while calculating, since the abacus also presents numbers as a collection of singularities. Empty sets, i.e. absent elements, are not countable. Consequently the zero is unknown in Greek and Roman cultures. The reason for this is that their numerical language is not constructed as a calculus; ‘calculus’ understood here as a written sign-system with the double function of being both the medium of numerical presentation and the instrument of numerical calculation. The decimal position system was developed in India, where a sign for zero was not only introduced in order to clarify numerical expressions, but where zero was used in calculations, and where rules were also developed for these calculations. The operative impetus to the decimal position system, also in so far as this only served a numerical repre-
11
From Zero to Infinity. What Makes Numbers Interesting. San Francisco: Mathematical Association of America, 1992. 1. Reid also distinguishes between zero as a ‘number’ and as a ‘symbol’. Ibid. 6.
466
Sybille Krämer
Fig. 1: Roman abacus and decimal notation.
sentation, can be clarified by a simple consideration. Independent of the historical proof of such a connection,12 the decimal position system can be at least systematically explained as the ‘writing down’ or transferral of the abacus technique into the medium of numeral-writing where the columns of the abacus remaining empty are marked in the numeral-configuration with ‘0’ (fig. 1). There is, however, also an etymological connection:13 ‘Sunya’ in Indian means ‘emptiness’ and was called ‘Sifr’ by the Arabs, which then as ‘Cifra’/’cipher (figure)’ at first only described the sign ‘0’, but then the whole numeral system. Does the word ‘emptiness’ play on the missing calculating stones in the corresponding column of an abacus? 12 13
A historical connection is unlikely. On the “systematic descent” of the Indian decimal system from the abacus, cf. ibid. 2f.; Rotman. Signifying Nothing. 10ff. Cf. Joseph Needham. Science and Civilization in China. 7 vols. Cambridge: Cambridge University Press, 1959. Vol. 3, 75.
The Productivity of Blanks
467
In any case, the trick of the decimal position system consists in being able to carry out the functions of the abacus as pure sign or writing operations, as calculations on paper. And with these calculations it is inevitable that the gap-sign also invades the function of a numerical sign, as soon as for example ‘4 – 4 = 0’ is calculated. The operative certainty inherent in algorithmic calculation prepares us gradually for the interpretation of ‘0’ as a ‘positive’ number. Nevertheless, the denial of nothingness so deeply rooted in Greek as well as Christian tradition stood in opposition to this. It remains a matter of speculation whether the association, widespread in Indian culture, of emptiness with something that, though it is ‘still’ nothing, can become something (an understanding of nothing as a potential and proneness) inspired the use of the mathematical zero. In the context of Western thought, however, the resistance to emptiness, the zero, and the vacuum was huge.14 It is not by chance that the struggle for an experimental proof of the vacuum,15 and for the mathematical recognition of zero as ‘nothing’ that is still ‘something’ go hand in hand within European scientific development. The cultural struggle between ‘abacists’ and ‘algorists’ nevertheless spanned centuries; and it was decided less by numerical-theoretical debates, than by the pragmatic demands of soaring trade-capitalism, which knew how to make use of the calculating superiority of the Indian numerals over the Roman numerical sign-system. Seen in mathematical terms, two ‘threshold phenomena’ are significant for the ‘0’ to receive the honorary status of a number. Firstly, the Analytical Geometry developed by Descartes where zero is placed at the center of the coordinate system.16 Ever since the Greek discovery of the incommensurability of the side and the diagonal of a square, multitudo and magnitudo, the countable and measurable, arithmetic and geometry, were treated as mutually non-translatable measures. To the extent that Descartes, with the help of the coordinate system, depicted points on number pairs, he showed the mutual translatability of geometrical figures and arithmetical numbers. Familiar with Indian numeral 14 15
16
Very interesting in relation to this: Robert Kaplan. The Nothing that Is. A Natural History of Zero. London: Penguin, 2001. 102ff. Cf. Hartmut Böhme. “Das Unsichtbare. Mediengeschichtliche Annäherungen an ein Problem neuzeitlicher Wissenschaft.” Performativität und Medialität. Ed. Sybille Krämer. Munich: Fink, 2004. 215-46; Wim Klever, ed. Die Schwere der Luft in der Diskussion des 17. Jahrhunderts. Wiesbaden: Harrassowitz, 1997. René Descartes. Geometrie. Ed. Ludwig Schlesinger. Darmstadt: Wissenschaftliche Buchgesellschaft, 1981.We neglect the fact that, for Descartes, the coordinates were not right-angled, and that he did not include the negative numbers in his system of coordinates. This was done by his followers.
468
Sybille Krämer
Fig. 2: Counting and Measuring.
calculation, Descartes was certain that the coordinate axes, to facilitate this translatability, could not begin with ‘1.’ In this way he avoided a mistake that our calendar still makes: if the birth of Christ is positioned at year 1, then the zero as caesura in the transition from counting the years from ‘before Christ’ to ‘after Christ’ is simply missing.17 How the translatability of the measurable into the countable is made possible by the introduction of zero is made clear by a simple thought experiment (fig. 2). If we put four sticks next to each other, mark the sticks 1, 2, 3, and 4, and from the first to the last, put each a step apart, then we need exactly three steps to reach the last, the fourth stick. We count four sticks and measure three steps. As soon as we begin the marking of the sticks with 0, so that the last has the figure 3, the quantity of the sticks and the quantity of steps is homogenized. 17
Cf. Dick Teresi. “Zero.” The Atlantic Monthly 280/1 (1997): 88-94.
The Productivity of Blanks
469
With this thought experiment something else becomes clear: the zero represents the point of view, the ‘starting point’ of the measuring subject. We will come back to this. Now we want to turn to the second mathematical threshold-phenomenon for the transformation of the gapsign into a number, and that is algebra. The early history of algebra knew an abundance of particular procedures for solving equations. However there was still no universal method which might, in a general way, express an algorithmic procedure for whole classes of similar equations. Two far-reaching innovations cleared the way for this. François Viète invented symbolic algebra, in which he not only denoted the unknown coefficients of an equation by letters of the alphabet, which was aleady in practice, but also the known. Now it was possible to write down the rules for the conversion of equations with the help of variables in a general way. Like the zero in a numeral configuration, the variables are also ‘blanks’ and ‘placeholders.’ However, zero was decisive for the progress of algebra in an even more direct sense. Again the problem to be solved is the universalization of particular algebraic procedures; though now it is no longer the case of the invention of a new ‘numerical language’ in the form of the alphabeticwriting of symbolic algebra, but rather an algorithmic procedure that consists of solving equations using zero. John Napier succeeded in homogenizing various particular methods by proposing ‘equations with nothing’ (fig. 3).18 Let’s stop here. The circle of the mathematical invention, in which the zero plays a meaningful role, does not reduce itself to analytical geometry and algebra. We cannot, however, step out of or get to the bottom of this circle. Instead, we can ask what is shown by these few contexts, sketched by us, where zero plays a role. 7. How Zero as Medium Makes the Heterogeneous Comparable We are interested in the connection between the appearance of the figure ‘0’ as an indispensable element of a decimal mode of writing numbers, and the emergence of the number zero as a mathematical object. Our emphasis on zero has at least made this much clear: zero is a number as soon as the figure ‘0’ is needed in mathematical expressions to represent a number. Therefore with the number zero, an epistemological object appears whose existence as a referent of the figure ‘0’ is due 18
Kaplan. The Nothing that Is. 140ff.
Sybille Krämer
470
a) al-Hwarizmis equation, since 825 in countless textbooks x2 – 39 + 8x = -2x x2 + 8x + 2x = 39 x2 + 10x = 39 There were three particular methods x2 + px = q x2 + q = px px + q = x2 b) Equation with zero: John Napier (1550-1617) Now universal method x2 + 10x – 39 = 0 x2 + 10x – 39 = (x – 3) (x + 13) according to: if ab = 0; then a, b or both = 0 x – 3 = 0 or x + 13 = 0 x=3 x = -13 Fig. 3: Algebra.
to the cultural technique of written calculation. Thinking further along these lines, we could, in good post-structuralist manner, discuss how, through movements of the signifier (the figure ‘0’), the signified (the number zero) is first of all crystallized; and how the materiality of a symbol, which can be used palpably, has a decisive effect on its semanticity. In short, how the use of a sign first of all creates its meaning (also in the referential sense of a relation to an object) can be demonstrated here almost prototypically. With such arguments we remain bound (as, incidentally, Brian Rotman is in his stimulating thoughts on zero19) to a semiotic perspective where anything interesting to be said about zero takes place in the exchange-relationship between sign and referent, and consequently obeys the semiotic regime, and the logic of the sign. However, it is precisely the anchoring of our considerations concerning zero within cultural techniques that has shown that what the figure ‘0’ does, is not only – though, naturally, also – the creation of a 19
Rotman. Signifying Nothing.
The Productivity of Blanks
471
new object, but also the provision, perhaps even ‘creation,’ of a flexibility for the calculating and mathematical act. If we examine closely what this flexibility consists of, it becomes clear that the operative spaces opened up by zero make it possible to connect formally separate or various entities and to homogenize them, and consequently to set free new kinds of calculating and mathematical practices. If this view can be confirmed, then the figure ‘0’ appears less as a sign-carrier, with the peculiarity of preceding by centuries its ‘belated’ referent, but rather as a medium; ‘medium’ understood here as a ‘middle’ and ‘mediator’ which makes heterogeneous worlds relatable, and mutually translatable. The supposition is therefore that the revolutionary potential of zero consists in its function of becoming productive as a medium. And that this mediatechnical potential distinguishes the zero from the other numerical signs of the decimal system – the zero fulfils a function here which cannot just as well be realized by the figures ‘1,’ ‘2,’ ‘3,’ . . . – although the zero can only develop its efficacy as an element of decimal calculation. To verify this thesis would require a thorough investigation which cannot be carried out here. We can only suggest in a brief outline why we think a media-view provides an interesting potential elucidation. The media-concept, which we consequently ‘bring into the discourse,’ can only be supposed, not deduced. Media mediate between heterogeneous fields or systems or worlds by enabling a transferral or exchange between two divergent sides, and as a result open up new space for cultural practices. Media can do this by embodying, by hybridization, attributes from both sides of that which is to be mediated. In this way they provide a connection between diverse things without having to surpass and annul this diversity. And now back to zero. The role of the figure ‘0’ as gap-sign in the decimal position system can be reconstructed in such a way – we have already demonstrated the beginnings of this – that the ‘empty place’ (leere Stelle) in the abacus is transferred into a ‘blank’ (Leerstelle) within a numeral configuration. For this transferral, two things are significant. While on the abacus, in a particular column, there are no stones, in the numeral configuration the missing and absence of something is marked with a sensibly perceivable sign. Nothingness is quantified;20 the emptiness is aistheticized. Exactly this possibility of making absence not only sensibly present, but also quantifiable and materially 20
On the role of the quantification as the cardinal attribute of modern European thought, cf. Alfred W. Crosby. The Measure of Reality. Quantification and Western Society 1250-1600. Cambridge: Cambridge University Press, 1997.
472
Sybille Krämer
usable, that shows that not merely does a blank on the abacus wander into a writing configuration, but also that a technical function (‘calculate with calculating stones’) is transferred to a symbol system (‘operate with signs’), with the performative effect that now the presentation of numbers is also a production of numerical relations; since calculation is nothing but this. The systematic connection between abacus technique and decimal calculation shows that the mediating function of zero consists in creating a hybrid which simultaneously makes technical and symbolic functions possible, and, with this pairing of technology and language or writing, in creating a medium that goes much further than the preceding calculative systems, as well as numerical systems. Or, let’s consider the coordinate system that becomes a medium for connecting geometry and arithmetic, the measurable and the countable. The zero is situated at the point of the intersection of the axes. This position is normally characterized by the zero being happily designated as the origin of the number lines that cross in it. Without wanting to discredit this origin-idea, we might characterise the place of zero less as a beginning, but rather as a middle, and a middle-point. By virtue of this placement the zero can set in motion an exchange between the space of the negative and positive numbers, between the present and absent, it makes a transition from the axis of positive value to the axis of negative value possible. Zero provides the connection between the y-axis and xaxis, and consequently can translate geometric points into combinations of the values of both axes. In any case, with this ‘reifying’ mode of what the zero does, we must constantly keep in mind that it is human or mechanical computers that put this, with the help of zero as the medium, to work. What the hybrid character of zero means here becomes clear with the example of the coordinate system. Zero is as much the beginning of the positive numbers on the x-axis as the negative numbers on the x-axis; it is the beginning of the positive and the negative y-values. What this comes to is not simply that it is a beginning, but rather that zero embodies the beginnings of two axes at the same time and in one. This ‘in one’ cannot be valued highly enough. By this kind of hybridization of two sides, which in a certain way neutralize each other, a kind of indifferentiality of the zero is achieved, the kind of neutrality, incidentally, which is concerned with maintaining the operational logic of each medium.21 In this way, zero clarifies what it means when, to make distinctions, we continuously require a medium. The zero marks a line 21
Zero is therefore not simply nothing, but according to 4 – 4 = 0, what results when two inverse procedures are used.
The Productivity of Blanks
473
of division that enables the distinction between two heterogeneous areas, to the extent that, in the medium, the zero can bring about the different and the common between what are distinguished. By ‘sitting’ in the middle of the number-series which divide the negative and the positive numbers, the zero also shows that, with the serious difference between having and not having, between presence and absence, between negative and positive in the homogenous countability provided by the zero, those things which diverge from each other are comparable and can be related, a countability and comparability, which now, for example, allows the question: what is the result of ‘seven subtracted from five’? It is the appearance of zero that makes this transition possible at all. 8. The Central-Perspectival Vanishing Point as Visual Zero Let’s turn now to the background of the above-mentioned media-theoretical perspective on the early modern period which – and actually significantly earlier than the zero became meaningful in the mathematicizing sciences – made the zero-point the increment of its central perspective pictures. This is the case of the vanishing point, that point within a picture constructed using central perspective, which is infinitely distant from the observer.22 In a picture not governed by perspective, the pictorial space is an associative space of objects whose mutual relationship is best defined by the meaning which these respectively embody: the important things are depicted as big, and those second in rank are depicted as smaller. By means of the vanishing point, the pictorial space is transformed into a system organized by coordinates in which the proportion of each object is calculable. The nearer the objects in the picture are to the vanishing point, the smaller they are in scale, and the further they are away, in the ‘fictive reality’ of the picture, from the viewer. It is no longer the meaning, but the relation between the ‘reality’ presented in the picture, and the viewing-process of the viewer, which produces the matrix of the coordinate system. What the vanishing point means can partly be understood in the way that it shifts the viewer into the picture as organizing center. These associations are familiar enough. Therefore we want to concentrate here on the question of what is actually gained as soon as the vanishing point is understood as a ‘visual zero.’23 22
23
Cf. Rotman. Signifying Nothing. 14ff.; Crosby. The Measure of Reality. 165ff.; Charles Seife. Zero. The Biography of a Dangerous Idea. New York: Viking, 2000. 95ff. Rotman. Signifying Nothing. 19.
474
Sybille Krämer
To begin with, there are certain basic affinities between the point and the zero: points are zero-dimensional objects. What that means, we can clarify operatively by imagining successively robbing three-dimensional objects of their dimensions.24 First a three-dimensional object is pressed flat so that it mutates into a two-dimensional plane, consisting of length and breadth. Then this flat structure is put on one of its sides, and pressed flat again, so that only a one-dimensional line exists. If this is pressed flat lengthwise, a zero-dimensional point results, which is without extension, that is, without height, breadth, or length. This is a characteristic of all points. What, however, characterizes the special relationship between the vanishing point and zero? Here too, an operative consideration can be introduced; though in this case not a thought-experiment but an actual experiment that the architect and sculptor Fillipo Brunelleschi staged in 1425. With this experiment, Brunelleschi wanted to document the perfect illusion of the central perspective picture, i.e. to show that our natural perception of world-objects, and the appearance of pictorial objects painted in central perspective, coincide. Placed in a row are: the observer, a picture of the baptistry in Florence turned towards the viewer, a mirror, and the baptistry itself. Brunelleschi drilled a hole in the picture so that an eye of the observer looking through this hole sees the reflection of the pictorial baptistry in front of the real baptistry, and in this way notes that he cannot distinguish between picture and reality. On the mirror surface, the usually invisible vanishing point of the picture is represented or visualized here by the eye of the observer. In 1435/36, Leon Battista Alberti theorized this experiment. He showed that a correct perspectival image could be constructed with a single vanishing point, which is placed in relation to a visual pyramid, whose tip lies in the eye of the viewer, and whose base in the object. The central-perspectival picture could then be defined as an even section through the visual pyramid.25 We analyzed the numerical zero in its quality of being a medium that provides a comparability between heterogeneous mathematical domains, and in this way enables ‘transitions’ between them. In analogy to this, we can interpret the visual zero of the vanishing point as a mediator between the real space and the pictorial space, between the space of objects and the projection space, between the observing subject and the observed object, between the I and the world. The vanishing point structures the picture like a coordinate system, whose zero-point it is. It 24 25
Cf. Seife. Zero. 96. Leon Battista Alberti. On Painting. New York: Penguin, 1991.
The Productivity of Blanks
475
is not by chance that part of the elementary exercises of perspective construction are tile patterns, reminiscent of coordinates, which in pictorial space undergo a mathematically exact transformation, and precisely for this reason appear as an illusionist development of the natural floor on which the viewer stands. In relation to what is depicted in the picture, mediated by the vanishing point, bodies and the intervals between them can now be handled in a unified way; body and non-body are, despite their qualitative difference, calculable with measure and number. Their size in the picture continuously decreases as it goes backwards, meaning in the direction of the vanishing point, whether it is a case of the objects or their interstices. Just as various axes come together by crossing each other at the numerical zero at the center of the analytical coordinate system, the projection lines meet at the visual zero. But not only that; above all, the organizing principle of all visible objects – which, like the vanishing point itself, is invisible – coincides with the position of the eye of the observer – and precisely this brings Brunelleschi’s mirror to mind. Brian Rotman imagines in the vanishing point the “visual equivalent of a demonstrative pronoun.”26 This is how a subject standing in a totally defined place, now and here, sees the scene. Embodied in the vanishing point, the subject is in the picture. Consequently, Jean Pélerin Viator has perceptively described the fixed point of the eye as ‘subject.’27 9. The Zero as Embodiment of the Modern Subject? Dirk Baecker establishes that at the zero, not only counting is reflected, but also the counters themselves, so that the history of the zero can also be understood as the history of the “hesitant discovery and the engaged denial of this observer.”28 In fact, not only in the case of the observer, but also for the subject is “the zero written on the body.”29 The numerical zero symbolizes the standpoint of the measuring and counting subject just as the visual zero marks the standpoint of the observing subject. In this way, an artistic construction principle anticipates what in 26 27
28
29
Rotman. Signifying Nothing. 19. Emanuel Alloa. Diaphanes. Vom Begriff zum Phänomen der ikonischen Differenz. Master’s thesis at the Institut für Philosophie of the FU Berlin, 2003. 19, who cites here Hubert Damisch. L’origine de la perspective. Paris: Flammarion, 1987. 14. Dirk Baecker. “Der Nullpunkt.” Preface to the German edition of Brian Rotman. Signifying Nothing: Die Null und das Nichts. Eine Semiotik des Nullpunkts. Berlin: Kadmos, 2000. 9. Ibid. 10.
476
Sybille Krämer
modern philosophy and science since Kant has been apostrophized as the ‘Copernican revolution’ – that we (only) experience the world as it shows itself from the subjective standpoint, meaning the perspective of the single viewer. The world of objects appears as a projection of the order of the subject. But is the role of zero actually appropriate for such a constitutivetheoretical, or at least constructivist-oriented view, of humanity’s place as subject of the world? We have seen that we can consider zero from a semiotic and medial point of view. Let’s try once again to focus the difference of both modes of viewing. Semiotically, the figure ‘0’ appears as a gap-sign in the position system, whose calculating function finally leads to the crystallizing of a new mathematical object: the number zero. The nothingness of emptiness becomes with this definition a something, which is also operable on paper, and, in the same way as the other numerical signs, thus prepares the ground from which the number zero can eventually grow. However, what is here considered in a prototypical way, using the figure ‘0’, is also the case for the other numberdefinitions. Not only the absence of a specific number, but rather all numbers are entities which are per se invisible, and which can only be included in the regime of visibility by means of numerical signs. That the visualization of something invisible in its cultural-technical use leads to the visualized arising as a mathematically recognized object, is a ‘fate’ affecting all numbers in the early modern period, which are no longer, as in Euclid’s time, treated as countable quantities of units, but rather as something that can be introduced as a referent of a calculated arithmetical expression. The signs create the objects that they name. From this semiotic perspective a performative constitution of mathematical objects by the use of a formal language can be clearly shown; and the debate on linguistic performativity is enriched with an important and previously neglected aspect. With this creation of an epistemological object using the culturaltechnical practices of sign-use, the epistemological role of the subject as creator of the world, which they can experience and investigate, seems to be confirmed once more. But doesn’t this connection show itself in another light when we adopt the media-perspective? From this media-perspective, the new, which is bound with the figure ‘0’, consists less in the creation of a new object. For media mediate between diverging domains by making them – precisely in their differentiation and despite their differences – comparable, and therefore opening up space for action whose productivity is rooted in interdependencies and new types of transaction between these domains. Before this ho-
The Productivity of Blanks
477
rizon, the figure ‘0’ emerged as that which mediates between calculating technology and numerical language, consequently bringing about written calculation, or which mediates between point and number, and so creating analytical geometry with its potential to solve geometrical problems algebraically. Brian Rotman has analyzed the zero semiotically and interpreted it as a meta-sign, a sign about a sign, which creates from the “emergence of a meta-subject,”30 an ‘original subject.’ Here, two different genitives of ‘subject’ are quoted, whose operative power consists of creating its sign-systems, and consequently also of being responsible for the emergence of the object which is signified by them.31 The enlightenment, which this semiotic perspective conceals, lies in a critique of the opinion that the ontology of objects precedes the construction of signs for these objects. This critique is fruitful, though it also draws on the traditional pattern of the subject as creator of the world. But what happens if in a media-theoretical orientation we understand zero as a mediator between diverging worlds, rather than as a mechanism for creating new worlds? What consequences would this have for a conception of the subject, since for everybody who is engaged with the zero, it is at least clear that zero is associable with the standpoint of the counting, measuring, observing subject? Could it be then that a new light would fall on that which defines human activities? Activities, whose ingenuity would not simply consist of making new things, but of offering new connections, from whose power new types of communicative and cognitive operations could emerge? With these questions, this text ends, which has nothing more in view than to show ‘family resemblances’ between science and art, using the example of zero.
30 31
Rotman. Signifying Nothing. 27. Ibid. 93.
478
Sybille Krämer
WORKS CITED Alberti, Leon Battista. On Painting. New York: Penguin, 1991. Alloa, Emanuel. Diaphanes. Vom Begriff zum Phänomen der ikonischen Differenz. Master’s thesis at the Institut für Philosophie of the Freie Universität, Berlin, 2003. Austin, John L. How to Do Things with Words. Cambridge: Harvard University Press, 2005. Baecker, Dirk. “Der Nullpunkt.” Brian Rotman. Die Null und das Nichts. Eine Semiotik des Nullpunkts. Berlin: Kadmos, 2000. 7-18. Böhme, Hartmut. “Das Unsichtbare. Mediengeschichtliche Annäherungen an ein Problem neuzeitlicher Wissenschaft.” Performativität und Medialität. Ed. Sybille Krämer. Munich: Fink, 2004. 215-46. Crosby, Alfred W. The Measure of Reality. Quantification and Western Society 12501600. Cambridge: Cambridge University Press, 1997. Damisch, Hubert. L’origine de la perspective. Paris: Flammarion, 1987. Descartes, René. Geometrie. Ed. Ludwig Schlesinger. Darmstadt: Wissenschaftliche Buchgesellschaft, 1981. Kaplan, Robert. The Nothing that Is. A Natural History of Zero. London: Penguin, 2001. Klever, Wim, ed. Die Schwere der Luft in der Diskussion des 17. Jahrhunderts. Wiesbaden: Harrassowitz, 1997. Krämer, Sybille. Berechenbare Vernunft. Kalkül und Rationalismus im 17. Jahrhundert. Berlin and New York: de Gruyter, 1991. Krämer, Sybille. “‘Schriftbildlichkeit’ oder: Über eine (fast) vergessene Dimension der Schrift.” Bild, Schrift, Zahl. Ed. idem and Horst Bredekamp. Munich: Fink, 2003. 157-76. Krämer, Sybille. “Die Rationalisierung der Visualität und die Visualisierung der Ratio. Zentralperspektive und Kalkül als Kulturtechniken des ‘geistigen Auges.’” Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst. Ed. Helmar Schramm et al. Berlin: Dahlem University Press, 2003. 50-67. Needham, Joseph. Science and Civilization in China. 7 vols. Cambridge: Cambridge University Press, 1959ff. Reid, Constance. From Zero to Infinity. What Makes Numbers Interesting. San Francisco: Mathematical Association of America, 1992. Rotman, Brian. Signifying Nothing. Basingstoke: McMillan, 1987. Schneider, Irmela. “Von der Vielsprachigkeit zur ‘Kunst der Hybridation.’ Diskurse des Hybriden.” Hybridkultur. Medien, Netze, Künste. Ed. idem and Christian W. Thomsen. Cologne: Wienand, 1997. 13-66. Searle, John. The Construction of Social Reality. New York: Free Press, 1995. Seife, Charles. Zero. The Biography of a Dangerous Idea. New York: Viking, 2000. Steiner, Reinhard. “Malerei als Spekulation.” Digitaler Schein. Ästhetik der elektronischen Medien. Ed. Florian Rötzer. Frankfurt a.M.: Suhrkamp, 1991. 435-54. Teresi, Dick. “Zero.” The Atlantic Monthly 280/1 (1997): 88-94. Vasari, Giorgio. The Lives of the Artists. Trans. Julia Conway Bondanella and Peter Bondanella. Oxford: Oxford University Press, 1998. Wirth, Uwe, ed. Performanz. Zwischen Sprachphilosophie und Kulturwissenschaften. Frankfurt a.M.: Suhrkamp, 2002.
JÖRG JOCHEN BERNS
Instrumental Sound and Ruling Spaces of Resonance in the Early Modern Period: On the Acoustic Setting of the Princely potestas Claims within a Ceremonial Frame* Ceremonial, which should ensure the frictionless traffic of different monarchic rulers among themselves as well as between ruler and court, and finally between ruler and subject, is known to be a complex order of signs that integrates all areas of the senses. Ceremonial functions not only as a choreography of the higher nobility and the princely family, but as an aesthetic specific to its formation, as “courtly aesthetic” per se, which calls upon all the senses and is set up to courtify, police, and discipline the broadest possible social sphere. To achieve this, instruments are needed to sharpen and strengthen the senses, extending their reach. From the descriptions and theory of ceremonial, as codified in Germany at the beginning of the eighteenth century in voluminous handbooks with collections of cases, and systematized in juristic and philosophical treatises, it would be possible to obtain a hierarchical scale or even a cooperative model of the instruments that are necessary for the production of signs and the maintenance of the order of signs. Since ceremonial theory offers not only a theory of behavior that revolves around a courtly decorum, but also contains – as a media-strategic system for the creation and maintenance of princely-courtly representation and pretension to power – a special instrumentology. As this immanent instrumentology has never – not even by early modern ceremonial theoreticians – been systematically revealed, one cannot expect the following reflections to fill this gap. They should, however, highlight the need for further research by isolating one aspect of this ceremonial instrumen*
Many thanks to Volker Bauer (HAB Wolfenbüttel), Bernhard Jahn (Universität Magdeburg), and Ulrich Schütte (Univsität Marburg) for their encouragement and editorial suggestions. JJB.
480
Jörg Jochen Berns
tarium (as much as this is possible): that of acoustic sign production by particular instruments that cannot easily or totally be assigned to the category of “musical instrument.” In these reflections, I refer, terminologically and empirically, not to books specializing in musicology or physics, but to the courtly ceremonial texts and festival diaries of the sixteenth to eighteenth centuries. This is because I am concerned with answering the question of what spatialities are opened by what sound instruments, or more precisely, what spatial types are made and served by what sound sources and what instrumentally or extra-instrumentally produced types of sound. The question of the princely court spaces of the early modern period limits, heuristically and systematically, the field of examples (and the field of the textual and visual sources to be interrogated) in advance. This is because, as is well known, the princely ceremonial is a system sui generis. According to medieval and early modern conceptions, every earthly ceremony is the realization of a heavenly ceremonial order that has descended to earth.1 This is the case for all social formations whether defined as clerical, secular, or, in a narrower sense, courtly. However, because every earthly ceremony has to be realized by means of all five senses2 – whether employed in isolation or synaesthetically – it must also be realized acoustically. In the early modern period, two heavenly models of ceremonial sound regulation could still be recalled: the pagan-mythological model of the music of the spheres3 and the biblical model of the heavenly choir as described in the vision of Isaiah.4 Both models can be combined in the 1
2
3
4
Here I offer an early modern variation on a thesis by Reinhold Hammerstein that is full of insights for the medieval period, developed by the most important ceremonial researcher arguing from the perspective of music history. Cf. Reinhold Hammerstein. Macht und Klang. Tönende Automaten als Realität und Fiktion in der alten und mittelalterlichen Welt. Bern: Francke, 1986. Cf. also the dissertation of the historian Sabine Zak. Musik als “Ehr und Zier” im mittelalterlichen Reich. Studien zur Musik im höfischen Leben, Recht und Zeremoniell. Neuss: Päffgen, 1979. On the aesthetic historical meaning of ceremonial in general, cf. Jörg Jochen Berns and Thomas Rahn. “Zeremoniell und Ästhetik.” Zeremoniell als höfische Ästhetik in Spätmittelalter und Früher Neuzeit (= Frühe Neuzeit, vol. 25). Ed. idem. Tübingen: Niemeyer, 1995. 650-66. As is generally known, the Pythagoreans had already developed the theory of the music of the spheres by transferring the laws of interval and tonal division that they had discovered to the planets. In their regular concentric movement around the center of the universe, tones would be produced analogous to those of the musical scale. Cf. Günther Wille. Musica romana. Musik im Leben der Römer. Amsterdam: Schippers, 1967, passim. Isaiah 6,1-4. Cf. also Ezekiel 1,24f. and 3,12f. For a detailed study of the medieval
Instrumental Sound and Ruling Spaces of Resonance
481
ruling ceremonial because, above all, both imply a constant harmonious sonority.5 The Pythagorean theory of universal harmony states that the cosmos is a harmonious structure conforming to natural laws that can be experienced musically, that identical laws operate in nature, in human hearing, and in music.6 According to Christian theologians, on the other hand, the heavenly music described by Isaiah is present in liturgical song and in the movement of the human pulse.7 Accordingly, both could be imitated in earthly ceremonial sign systems. The harmony of the heavens prevails in the harmonics of human music since, thanks to mathematics, heavenly harmonics can be understood by humans, and corresponds with human interval theory, which is also effective in the shape and measure of human musical instruments. Therefore it can be claimed that without ceremonials there would be no music. This sentence, however, is not reversible. If it can rightly be claimed from a historic and genetic point of view that music arises out of the ceremonial – the earthly and the heavenly – then the acoustic signaling of earthly ceremonials is not identical with the musical sphere. Rather, the acoustic dimension of court and state ceremonial clearly goes beyond music
5 6
7
reception of this “heavenly liturgy,” cf. Reinhold Hammerstein. Die Musik der Engel. Untersuchungen zur Musikanschauung des Mittelalters. Bern and Munich: Francke, 1962. 17ff. Hammerstein. Musik der Engel. 116ff. shows in detail the theological-historical steps that ensure the transferral from a “world sound” to a “world liturgy.” The fundamental importance still accorded to the theory of a universal harmony in the natural sciences of the seventeenth century can be studied in such varied texts as Johann Kepler’s Mysterium cosmographicum (1596) (cf. on this Rudolf Haase. Johann Keplers Weltharmonik. Der Mensch im Geflecht von Musik, Mathematik und Astronomie. Munich: Diederichs, 1998), Marin Mersenne’s Harmonie universelle (1636), or Athanasius Kircher’s Musurgia universalis (1650). Kircher’s influence can also be seen in the diverse editions, translations, and reworkings of his musicological œuvre, e.g. in Artis magnae de consono & dißono ARS MINOR; Das ist/ Philosophischer Extract und Auszug/ aus deß Welt-berühmten Teutschen Jesuitens Athanasii Kircheri . . . MUSURGIA UNIVERSALI . . . Ausgezogen und verfertiget . . . von Andrea Hirschen/ . . . Evangel. Pfarrern zu Bächlingen . . . Leipzig: Zentralantiquariat der DDR, 1988 [facsimile of the edition Schwäbisch Hall, 1662] or in ATHANASII KIRCHERS è Soc: JESU Neue Hall- vnd Thon-Kunst/ Oder Mechanische Gehaim-Verbindung der Kunst und Natur . . . Jn unsere Teutsche Mutter-Sprach übersetzet von AGATHO CARIONE. Hanover: “Libri rari” Schäfer, 1984 [facsimile of the edition Nördlingen, 1684]. Cf. Werner Kümmel. “Puls und Musik (16.-18. Jahrhundert).” Medizinhistorisches Journal 3 (1963): 269-93; as well as Jörg Jochen Berns. “‘Vergleichung eines Vhrwercks, vnd eines frommen andächtigen Menschens.’ Zum Verhältnis von Mystik und Mechanik bei Spee.” Friedrich von Spee. Dichter, Theologe und Bekämpfer der Hexenprozesse. Ed. Italo Michele Battafarano. Gardolo di Trento: Reverdito, 1988. 101-206.
482
Jörg Jochen Berns
and its instruments. The sound ceremonial not only makes use of music, but also operates in the foreground and at the edges of compositionally and instrumentally regulated music. It operates between the poles of contourless noise and absolute silence. This paper will discuss the courtly-ceremonial possibilities and laws that arise between the poles of noise and silence for the princeps absolutus of the early modern period under three headings. Firstly, different acoustic causes of the courtly ceremonial, thus different scenic soundcasus (I) will be introduced. Whereupon suggestions for a typology of resonance-spaces (II) of the courtly-absolutist world will be made. And finally, problems of an acoustic instrumentology (III) and the ceremonial signal-potential latent in different musical instruments and nonmusical devices will be sketched out. I. Sound-Casus The aspect of the sound-casus will be illustrated by five exemplary scenes. (A systematic understanding of all casus and their scenic variables could only be achieved through a complete evaluation of all early modern ceremonial books.8) Scene 1: The Empire has a New Leader.9 A white flag ascends the spire of Frankfurt Cathedral. The cathedral’s bells ring out at full peal. At the foot of the tower, from a specially installed platform, a specially designated proclamator informs the awaiting citizens of Frankfurt and the world that the Empire has a new impe8
9
The beginnings of a classification of courtly and state ceremonial-casus is offered by the compendia of Friedrich Wilhelm von Winterfeld, Gottfried Stieve, Johann Christian Lünig, Julius Bernhard von Rohr, Friedrich Carl Moser, and Johann Jacob Moser. Cf. also the classification of the Herzog August Bibliothek Wolfenbüttel. Wolfenbüttel Thesaurus for Early Modern Festival Books. Online source: http:// www.hab.de/forschung/projekte/festkultur.htm (July 2007). Cf. Johann Christian Lünig. Theatrum ceremoniale historico-politicum, oder Historisch- und Politischer Schau-Platz aller Ceremonien. Leipzig, 1719/20. Vol. I, 1138ff. as well as the dissertation of my student Bernd Herbert Wanger. Kaiserwahl und Krönung im Frankfurt des 17. Jahrhunderts. Darstellung anhand der zeitgenössischen Bild- und Schriftquellen und unter besonderer Berücksichtigung der Erhebung des Jahres 1612 (= Studien zur Frankfurter Geschichte, vol. 34). Frankfurt a.M.: Kramer, 1994.
Instrumental Sound and Ruling Spaces of Resonance
483
rial leader. The citizens burst into cheers. This cheering, supported by fanfares of trumpets and drum rolls, should be understood as an extension of the acclamatio, the confirmation of the new ruler by the people.10 Meanwhile, the peals of the election and coronation church are heard in the church towers of the surrounding districts of Frankfurt and sent on in ever-greater radii to the church towers of the towns and villages lying outside Frankfurt. The concentrically pulsing sound vacuole reaches to the periphery of the kingdom; “Another correlation thereby arose between the loudness of a bell and the extent of a parish or commune’s territory,” to which Alain Corbin has referred to in his essential work on the language of bells,11 was attempted by the territory of the Empire. If in the Middle Ages and early modern period, the reach of a bell commonly defined the extension of a parish, then the imperial ringing extending from the Frankfurt coronation church is intended, by a concentric wave-shaped linking of all the bells of the Empire, to define the whole imperial territory. After the bell ringing at the Frankfurt cathedral had begun, which was also greeted and amplified by cannon fire on the city’s ramparts, the city gates were opened. These were kept strictly closed during the election and coronation, converting the intramural space into a large conclave, in which the smaller conclave of the election chapel in the cathedral was protectively enclosed, frequently for weeks. Now however, after the proclamation and acclamation, and in response to the bell signal, they were opened again to allow the couriers to fly off in all directions to inform Imperial States (Reichsstände) such as Prague, Nuremberg, or Cologne. One can imagine these couriers riding in relay for hours through the sound-vacuoles emanating from the bells of Frankfurt and, in part, competing with the spreading waves of sound. This competition would combine two very disparate experiences of speed – the speed of riding and the speed of the bells being transmitted from tower to tower – so that the battle may well have seemed hopeless from the beginning. But then the couriers would arrive in a less populated region, deprived of bells and thus silent, and their efforts would become more meaningful again. When they finally reached their goal, namely sovereign cities such as Nuremberg, Prague, or Cologne, the staging of sound staging would 10
11
On the tradition of the ‘acclamatio,’ cf. G. Langgärtner and G. May. “Akklamation.” Lexikon des Mittelalters. Munich: Deutscher Taschenbuch Verlag, 2002. Vol. I, 213f.; Zak. Musik als “Ehr und Zier.” 9ff. On the organ as acclamation instrument, cf. ibid. 172ff. Alain Corbin. Village Bells. The Culture of the Senses in Nineteenth-Century French Countryside. Trans. Martin Thom. New York: Columbia University Press, 1998. 97.
484
Jörg Jochen Berns
be repeated after the Frankfurt model. Here too a clamor of the city population supported by fanfares and drums extended from the center, and the bell ringing of the central church was sent in concentric rings out to the neighboring churches and the town and village churches of the territory, and sustained for hours, while shots were regularly fired from the ramparts.12 Such media-specific staging of imperial or even princely election and coronation acts were of course not specific to the politics of the Empire. As the close intertwining of the whole European higher nobility and its ceremonial codes would lead one to expect, the acoustic ceremonial staged at Christian courts outside the Empire deviated only slightly from those of the Emperor and the upper nobility of the Imperial States. The same features of human shouts, trumpets and drums, bells, and cannons, as well as particular genres of music were present here, too. A contemporary report of the coronation of King George I in London in 1714 recalls: The cheering and jubilation of the people was so great that throughout the whole city a joyful vivat rang out accompanied by the constant ringing of bells and the firing of thunderous guns so that it seemed as if the joyful calls were being spread out so far through the distribution of the air that the jubilant cheering of the neighboring provinces could be united with the wishes of the capital. And since the day did not seem sufficient to provide enough for their demonstrations of joy, these were continued during the illuminations and bonfires and fireworks until late into the night.13
Scene 2: Peace Celebrations As with the election and coronation of a monarch, peace celebrations were also a casus where the whole population was affected and needed to be included.14 Therefore it is not surprising that the acoustic signaling of both casus have certain media-strategic similarities. Let us consider the celebrations to mark the Peace of Westphalia.15 In Osnabrück 12 13 14 15
Cf. Wanger. Kaiserwahl und Krönung. 142, as well as the coronation diaries quoted there. Anonymous. “Nachricht von der solennen Crönung König Georgens des I. in Groß-Britannien, de Anno 1714.” Lünig. Theatrum ceremoniale. Vol. I, 1384. Various peace celebrations are described in Lünig. Theatrum ceremoniale. Vol. I, 768ff. On this, cf. the three volume catalogue by Klaus Bußmann and Heinz Schilling, eds. 1648. Krieg und Frieden in Europa. 3 vols. Munich: Bruckmann, 1998. The source references and expositions on the acoustic side of the ‘solemnities’ connected with
Instrumental Sound and Ruling Spaces of Resonance
485
and Münster, the Peace of Westphalia could not be celebrated as a local or patriotic concern because they both contained too many regional and international parties. The peace could then be celebrated all the more magnificently and emphatically in imperial free cities such as Nuremberg16 or in princely seats such as Weimar – as a festival to which all strata of the population could relate. As described in a contemporary report on the celebrations in Weimar: Anno 1650, 19 August, Monday early at three o’clock the start of the . . . peace celebration took place in the princely seat of Weimar, 20 pieces as well as 4 mortars standing behind the princely castle, and partly filled with shells, fired away to the attention of our neighbors, where also, according to orders, all the bells started to ring throughout the whole principality, and then in all the towers of the towns and villages while trumpets, pipes, shawms, and other instruments were played, and people sang and praised God. The princely musicians waited by the royal seat, trumpeter and military drummer in the royal castle, or socalled Welsh garden above on the newly-built snail and lime houses with their instruments and made music, and blew at first an intrada, then alternately with one versicle following another, the psalm: O give thanks unto the Lord, etc. which lasted until almost five. After this the 4 mortars were fired again, and then the town choir and town pipers sang and blew in the town hall the psalm: Bless the Lord, O my soul, etc. and other pious songs and odes. At five all the bells were rung another time, whereupon the trumpets and military drums were blown and beated once more. At six all the bells were rung for the third time at court and in the city, and again trumpeters and drummers played; whereupon the court service started.17
Here too, a movement back and forth in the ceremonial total arrangement can be clearly discerned, an alternation and interrelation between local musical staging and spacious sound demonstration by differentiated instrumental ensembles and resounding coarse-toned instruments that acoustically overarch and politically define the whole territory.
16
17
the Peace of Westphalia, here vol. 2, 409ff., are – and this is characteristic of the generally limited interest in musical and acoustic ceremonial – scant; in any case, they do not even draw on Lünig’s contribution. An overview of the peace celebrations in Nuremberg is offered by Hartmut Laufhütte. “Das Friedensfest in Nürnberg 1650.” 1648. Krieg und Frieden in Europa. Ed. Bußmann and Schilling. Vol. 2, 347-58. On the acoustic-onomatopoeic dimensions of the Nuremberg poets, cf. Jörg Jochen Berns. “Kriegs- und Friedensbilder. Mittel ihrer ästhetischen Reflexion im 17. Jahrhundert.” Morgen-Glantz 9 (1999): 181-218. Anonymous. “Beschreibung des prächtigen Festins, so wegen des Westphälischen Friedens an dem Fürstl. Hofe zu Weymar gehalten worden, de Anno 1650.” Lünig. Theatrum ceremoniale. Vol. I, 828f.
486
Jörg Jochen Berns
Scene 3: The Prince is Dead and must be Buried. When death caught up with a prince far from his residence, the body was transported to his residence in convoy for embalming, presentation, and burial.18 Along the route, all the parish bells would ring in response. The deceased traveled through a tunnel of sound while his convoy left sound-trails consisting largely of rumbling drums.19 During the burial service in the court chapel, all the churches throughout the land would hold mourning services where simultaneously the same verses from the Bible were read out, sermons on the same texts were given, the same songs were intoned, and a constant pealing in all the parishes would bind the land together into a single acoustic space, sometimes for 24 hours.20 Scene 4: The Prince Conducts. Elias Canetti correctly noted: There is no more obvious expression of power than the performance of a conductor. Every detail of his public behavior throws light on the nature of power. Someone who knew nothing about power could discover all its attributes, one 18
19
20
On the pompe funebre in general, cf. Jill Bepler. “Ansichten eines Staatsbegräbnisses. Funeralwerke und Diarien als Quelle zeremonieller Praxis.” Zeremoniell als höfische Ästhetik. Ed. Berns and Rahn. 183-97. Despite being otherwise rich in facts, the dissertation by Rudolf J. Meyer. Königs- und Kaiserbegräbnisse im Spätmittelalter. Von Rudolf von Habsburg bis zu Friedrich III. (= Forschungen zur Kaiser- u. Papstgeschichte des Mittelalters, vol. 19). Cologne, Weimar, and Vienna: Böhlau, 2000, unfortunately pays no attention to the acoustic side of the burial ceremonial. Among the early modern ceremonial sources, particularly informative is Claude-François Menestrier. Des décorations funèbres. Paris, 1683, as well as Lünig. Theatrum ceremoniale. Vol. II, 552-705: “Vom Ceremoniel bey Beysetzungen, Leichen-Processionen, Begäng- und Begräbnissen.” Cf. also Jörg Jochen Berns et al., eds. Erdengötter. Fürst und Hofstaat in der Frühen Neuzeit im Spiegel von Marburger Bibliotheks- und Archivbeständen (= Schriften der UB Marburg, vol. 77). [Exhibit. cat.] Marburg: Universitätsbibliothek, 1997. 348-71. Rohr, in the chapter “Von Leich-Begängnissen und Begräbnissen” of his Einleitung zur Ceremoniel-Wissenschafft Der großen Herren writes: “§. 11. Should the prince’s body be brought from one place to another to its princely tomb . . . the bells are rung not only in all the villages through which the funeral procession passes, but also in all the surrounding villages as soon as there is a sign of the body from a distance . . . From the fortresses, they are accompanied by the firing of cannons.” Julius Bernhard von Rohr. Einleitung zur Ceremoniel-Wissenschafft Der großen Herren. Ed. Monika Schlechte. Leipzig, 1990 [facsimile of the edition Berlin, 1733]. 309f. On the various possible ways of creating sound-spaces by means of bells, cf. the detailed study by Corbin. Village Bells. Cf. also Zak. Musik als “Ehr und Zier.” 370ff., 108ff., and 133ff.
Instrumental Sound and Ruling Spaces of Resonance
487
after another, by careful observation of a conductor. The reason why this has never been done is obvious: the music that the conductor evokes is thought to be the only thing that counts; people take it for granted that they go to concerts to hear symphonies and no one is more convinced of this than the conductor himself. He believes that his business is to serve music and to interpret it faithfully.21
As it has been noted that, at the premiere of a concert or later also an opera, a number of princes – for instance Emperor Maximilian I, Anthony Ulrich, Duke of Braunschweig-Lüneburg, Emperor Leopold I, or Frederick II king of Prussia22 – marked the opening bars themselves by visual and/or acoustic signs by taking up the conductor’s baton (which only came into use at the turn of the eighteenth century, initially as a large beating staff)23 or even played an instrument, one could be satisfied with Canetti’s observation that there is no clearer expression of power than the performance of the conductor, and that the absolute ruler would have certainly known this or felt this. However, the genesis of the relationship prince/conductor should be understood differently. It was not the prince who imitated the conductor, but rather the conductor the prince. The post and profession of the conductor first emerged in the frame of the early modern principality and courtly society, partly because the musical formations – particularly the large symphony orchestra – that made such a position necessary in the first place only 21 22
23
Elias Canetti. Crowds and Power. Trans. Carol Stewart. New York: Continuum, 1973. 394. The degree to which Emperor Maximilian I insisted on his privilege of invention and staging is revealed in the program information that he added to the musical presentation of his ‘Triumphzug’ (triumphal procession) – a gigantic series of 147 woodcuts that established a new genre within German cultural space. Cf. Der Triumphzug Kaiser Maximilians I. 1516-1518. 147 Holzschnitte von Albrecht Altdorfer, Hans Burgkmair, Albrecht Dürer u.a. Mit dem von Kaiser Maximilian diktierten Programm (= Harenberg Kommunikation, vol. 100). Afterword by Horst Appuhn. Dortmund: Harenberg Kommunikation, 1979. Here, the leader of a musical ensemble of lute and viol players (Bl. 17/18) is said to have uttered the words: “The bold and resonant tone, I have nobly and beautifully produced, for great delight, by imperial example.” (175); similarly Maximilian allows the Musica-Meister of a group of shawm, trumpet, and crumhorn players (Bl. 19/20) to announce that he performs courtly music “as His Imperial Majesty had given the same to me” (176). And the “Capelmaister” Jörg Slakany, the leader of the “Musica Canterey” (Bl. 25/26) remarks: “How on the instructions of the Emperor he has created the sweetest order in the singing of the Canterey. Following the righteous method and concordance, also symphony and ordinance, juncture and some melody, I have improved the cantory, according to my disposition brought to me by the Emperor” (177). In the 1660s, Emperor Leopold I conducted operas himself before a small, illustrious public. Cf. Herbert Seifert. Die Oper am Wiener Kaiserhof im 17. Jahrhundert (= Wiener Veröffentlichungen zur Musikgeschichte, vol. 25). Tutzing: Schneider, 1985. 67, 322, and 406.
488
Jörg Jochen Berns
emerged in the culture of princely courts. If, after a few bars, the early modern princes would then pass the baton on to a professional Kapellmeister, who would continue where the former had left off, this only testifies to such a genesis. The musical première – which even today is still given a certain precedence – had an authority in the early modern period that was ensured by neither the musicians nor the public, but actually by the prince and his conducting. The initial conducting signals two things. Firstly, that the office of conductor is not a permanently conferred post, a stable occupation, but a temporarily bestowed and delegated princely prerogative – therefore a princely privilege that must be bestowed again with each performance. Accordingly, the conductor’s baton is only a specific variant of the princely scepter, comparable with other similarly conferred interpreting-instruments and commanding batons such as the marshal’s baton of the military supreme commander, the crosier of the bishop,24 the ceremonial staff of the steward (Hofmeister), or the scepter of the university rector. Secondly, the initial conducting of the Princeps signaled that he, in his own person, determined the beat of the music by the beat of his pulse. Before the introduction of the metronome in the late eighteenth century, all musical tempi were determined by the pulse of the conductor.25 The five main degrees of the musical tempi (largo, adagio, andante, allegro, presto) as well as the sub-degrees (vivace, moderate, grave, etc.) were always measured ad hoc in relation to the conductor’s pulse. Therefore when the prince conducted, the tempi of his pulse, his blood, his heart would be conveyed through the musicians’ performance to the audience. The prince transposes his pulse into the audible; his blood emanates sounds. The Weimar courtier Johann Wolfgang von Goethe, mindful of such traditions, did not hesitate to interpret the ceremonialized life at court in general as a concert: “Court life resembles a music where everyone must keep his timing and breaks,”26 as he remarked to Eckermann on 16 August 1824.
24
25 26
In the course of European history, the crosier was, of course, only rarely and increasingly rarely used by princes, but increasingly by cardinals and, indirectly or directly, by the pope. Cf. Kümmel. “Puls und Musik.” Johann Peter Eckermann. Gespräche mit Goethe in den letzten Jahren seines Lebens. Berlin: Aufbau-Verlag, 1982. 106.
Instrumental Sound and Ruling Spaces of Resonance
489
Scene 5: The Princely Eavesdropper Because God is all-seeing and all-hearing despite being invisible and silent, the prince must also endeavor to see and hear as much as possible without being heard and seen himself. The early modern period was the first to set up an extensive system of diplomacy as well as an intricate network of informers, including espionage, to function as a compliment to public representations and propaganda. Besides the human eavesdroppers, there was also an increasing interest in technical listening devices. Athanasius Kircher, who did not dedicate his teaching and instruction book Musurgia universalis27 to the princes of Europe merely by chance, reflected on how princely power could be secured acoustically – by subjecting the people to a barrage of sound, as well as by means of listening systems. The listening system is clearly not part of the ceremonial, but something that operates behind this. As a secret power, it has a negative relation to the demonstrative public nature of the ceremonial. Among other things, Kircher experimented with listening horns. In an engraving (fig. 1), a large shell-like device is depicted receiving various noises such as a conversation in the castle court or in a covered walk; theses sounds are then concentrated towards a hidden point in a hollow portrait bust where the eavesdropping prince can listen in secrecy. A listening complex where several listening funnels from several rooms are led together into the room of an eavesdropper is depicted in another print (fig. 2). Here it is already a matter – two and a half centuries before the invention of the microphone – of the installation of the kind of control room that is still installed – admittedly in a finer form – by secret services of every shade in buildings where decisions of a political nature are made. The fact that in the early modern period it was already evident that every receiving device is also potentially a transmitting device can be seen in another of Kircher’s prints (fig. 3). This shows a primitive form of the music broadcast – a music ensemble whose sound product is siphoned up by a ladle-shaped device and projected to the exterior, providing a place behind the castle, a garden space for instance, with musical entertainment. 27
Athanasius Kircher. Musurgia universalis sive ars magna consoni et dissoni in X. libros digesta. 2 vols. Rome, 1650. The work appeared in many editions and various translations and reworkings.
490
Jörg Jochen Berns
Fig. 1: The prince’s eavesdropper. Snail-shaped listening device in a castle building. Copperplate engraving from Athanasius Kircher. Musurgia universalis (Rome, 1650).
Fig. 2: Listening complex in a court building, consisting of three listening domes integrated into the ceiling whose acoustic results are achieved by means of funnel-shaped tubes in the eavesdropper’s room. Copperplate engraving from Athanasius Kircher. Musurgia universalis (Rome, 1650).
Instrumental Sound and Ruling Spaces of Resonance
491
Fig. 3: Baroque musical broadcast. Transmission of an instrumental concert from the inner courtyard into the open-air. Copperplate engraving from Athanasius Kircher. Phonurgia nova (Kempten, 1673).
Fig. 4: Baroque “radio broadcast.” Carrying sound to a territory by means of several horns installed on a hill. Copperplate engraving from Athanasius Kircher. Phonurgia nova (Kempten, 1673).
492
Jörg Jochen Berns
Finally, Kircher also considered how a whole landscape could be barraged with sound from a single point. Another print (fig. 4) shows a baroque prefiguring of radio broadcasting. All these inventions served the rulers’ control of sound. Like no other scholar of his period, Kircher politicized the acoustic. He knew that the art of sound requires specific instruments, and that this must be linked to an art of resonance with resonance and listening installations, so that both interactively space-creating and space-controlling could be made useful to an art of ruling. These observations are naturally also relevant to ceremonial praxis; above all when one considers that the ruling ceremonial must make use of different resonance-spaces, either already existing or yet to be created. II. Typology of Resonance-Spaces Ceremonial books and festival reports allow one to conclude that the ruling art of resonance distinguishes four important types of resonancespace: a) the immobile and invariable resonance-space, b) the mobile and internally variable resonance-space and c) the immobile and internally variable resonance-space. These three courtly-ceremonial and therefore sign-permeated resonance-spaces are all embraced as arbitrary, artificial spaces of d) the everyday-natural resonance-space (as space of contingent noises). The internal structure and function of these four resonance-spaces can be illustrated with examples: ad a) One example of an invariable immobile resonance-space, which is not merely used ceremonially but also conceived ceremonially is the church space with its immobile “fixtures,” the fixed points of sound distribution: altar, pulpit, organ. The princely throne room with its specific furnishing – with the throne in front of a heraldically marked wall, with canopy, throne steps and certain open spaces in particular relation to the doors – is another invariable immobile resonance-space. Here, sound is directed through the fixed architectural elements made from particular building materials as well as through textile additions (curtains, tapestries, carpets, etc.) in such a way that it remains calculable even in its potential for redundancy and reduplication – including the echo effects much loved by the courtly baroque.28 28
Cf. Jörg Jochen Berns. “Die Jagd auf die Nymphe Echo. Künstliche Echoeffekte in
Instrumental Sound and Ruling Spaces of Resonance
493
ad b) A mobile internally variable resonance-space, on the other hand, is a space created by the prince himself, which permanently surrounds him, and which accompanies him through all his movements: on journeys, while hunting, or at war. The prince builds around himself an acoustic cocoon. This is largely produced by trumpets and drums, while hunting, however, by horns, and on journeys occasionally by transportable organs.29 In 1612 for instance, Matthias, the pretender to the crown, rode into Frankfurt with 38 musicians, including 16 trumpeters, a drummer, and an organist.30 The early modern princes also occasionally had so-called musical coats of arms, namely sound-signals or signature tunes, which would signal their arrival. Such musical coats of arms were also used as motif in tribute pieces (Huldigungsmusiken) and even operas. ad c) An example of an immobile, but internally variable soundspace is the open ground in front of and behind the castle. For the sake of simplicity, I shall limit myself to the three-wing model as built in Versailles and then later throughout the Empire in innumerable variants. The baroque castle of this type has two fronts. Different communication interests prevail on the town-side front than on the park-side. The town-side front opens onto the traffic of the world administered by trade, offices, and the military, while the other front opens – mediated by garden and park – onto the open countryside. One can generalize in ideal terms: both sides open onto two worlds that are so fundamentally different that they seem unconnected, even unconnectible. The street or town-side with the cour d’honneur historically and strategically follows the situation of the transportation links and fortification. The park-side on the other hand is divided by the imperious gaze of the monarch into extensive channels – the imperious gaze that, wandering over carpets of formal flower-beds, gravel path, and water surfaces, passing through
29
30
Poesie, Musik und Architektur der Frühen Neuzeit.” Die Mechanik in den Künsten. Studien zur ästhetischen Bedeutung von Naturwissenschaft und Technologie. Ed. idem and Hanno Möbius. Marburg: Jonas Verlag, 1990. 67-82. When visiting towns and cities, Emperor Frederick III (1436-1476) loved to bring along a musical organ automaton in the shape of a powerful ox. Cf. Rudolf Quoika. Altösterreichische Hornwerke. Berlin: Merseburger, 1959. 65f. During his ‘triumphal procession,’ Frederick III’s son, Emperor Maximilian I, brought along his court organist, Paul Hofhaimer, in a carriage pulled by a dromedary, which was meant, according to information in an inventory program stemming from Maximilian himself, to make it conspicuous, “how he [i.e. Hofhaimer] modified the performance according to the Emperor’s instructions.” Der Triumphzug Kaiser Maximilians I. 176. Cf. also Hammerstein. Macht und Klang. 96. Wanger. Kaiserwahl und Krönung. 140.
494
Jörg Jochen Berns
green hedges and walls of boscage, can picture itself in the ideal and possibly even the allegorical. If on the park-side – inaccessible to the public, and not even visible – other communication interests prevailed than on the town-side, then a different acoustic-ceremonial code must have corresponded to this. But how was this structured? As can be shown, due to the immobility of the resonance-space, both sides included contingent noises but also artificially introduced sound elements. The town-side, from an acoustic point of view, is more realistic, while the park-side opens itself to the ideal. This is because on the town-side the prosaic noises of the everyday prevail, while on the park-side natural-musical and artificial sounds prevail – birdsong in the boscage as artificial woods, but also in specially created aviaries. And here there are a few sound installations and resonance-architectures: water organs, twittering machines, concert pavilions, open-air theater. Here, the desired sounds are also specially exhibited in sculptures. There are stone and brass tritons, fauns, satyrs, naiads, herdsman, hunters, and so forth, all supplied with emblematic and characteristic instruments. The park-side is the side of a perpetual divertissement31 where art and artificiality cooperate, while the townside presents itself more soberly as the side of business, of the military, of office and duty. It should be noted at this point that the French park, particularly in the prototypical grounds of Vaux-le-Vicomte and Versailles, is not exhausted in this acoustically diverting function. If it has been rightly suggested that its visual construction contains or exposes “the skeleton of the central perspective image,”32 and that it offers a ‘landscaping of the central perspective framework,’ then it should also be stressed that the acoustic arrangement of the park of Versailles knows and recognizes no central perspective. The acoustic divertissement opens itself only through a divergence from the visual central perspective model. The natural and artificial acoustic potential is revealed only by leaving the hall of mirrors of the palace to walk through the park. Every walk in the park is simultaneously an avoidance of its visual structure. The places of divertissement are only found next to the visual axes.
31
32
Among the courtly-princely “divertissement” that plays a not inconsiderable role in German ceremonial literature of the eighteenth century, Rohr includes all “festivities that flatter the outward senses,” to which they can bring “refreshment and amusement” to the great men worn out by the “burdens of ruling.” Rohr. Einleitung zur Ceremoniel-Wissenschafft Der großen Herren. 732ff. Martin Burckhardt. Metamorphosen von Raum und Zeit. Eine Geschichte der Wahrnehmung. Frankfurt a.M. and New York: Campus-Verlag, 1994. 188.
Instrumental Sound and Ruling Spaces of Resonance
495
It can generally be said about the function of the ruling spaces of resonance that whether the prince is lingering in the immobile resonancespace of the representational rooms of a castle or a church, whether he is en route in a mobile sound-cocoon, whether he receives visitors on the steps of the town-side entrance or is diverting himself in the park with guests, he never finds himself in a simple everyday acoustic. Just as the ceremonial shuns all accident, it also shuns accidental noises. It operates with a changing sound scenery that protects the prince from everyday-accidental noise pollution, or ennobles this by giving it a technically abstract form. Not only wild birds that are accidentally present are heard in the park of the castle, but also mechanical, artificial birds.33 Here, no river or brook babbles through the park as decreed by nature, but a water guided by canals, pipes, pumps, and sluices, whose acoustic qualities – the babbling and splashing of artificial waterfalls, the murmuring and trickling of artificial streams, the splashing of fountains – are mentioned in numerous plans and garden descriptions. III. On Instrumentology A theory of instruments is also a constitutive feature of acoustic ceremonial. The sounds created by means of various instruments carry different distances, they create various sound radii. At the same time, instrumental diversity presents a scale of tonal colors, acoustic values, which need to be psychosemantically distinguished because their potential to produce affect differs. Cannons, bells and organs, trumpets and drums, as well as the human voice have the potential to be projected over large distances as well as having a psychologically expressive quality that clearly differs from both the intimate acoustic articulations of stringed instruments and the finer woodwind and brass instruments. It is noticeable that the early modern prince used all – pre-technical as well as technical – instrumental possibilities, the crude as well as the fine, in contrast and combination, for his ceremonial purposes, occasionally going so far as operating as a composer himself. Because the prince must be 33
“The attraction of simulating birdsong lay in the comparison with the natural birdsong that could be heard near to the aviaries – there was hardly a pleasure park that did not hold a variety of indigenous and exotic birds. The automatons, and other mechanical installations staged the contest between art and nature that were fought out in the princely garden and in the grottos.” Birgit Franke. “Automaten in höfischen Lustgärten in der Frühen Neuzeit.” Automaten in Kunst und Literatur des Mittelalters und der Frühen Neuzeit (= Wolfenbütteler Mittelalter-Studien, vol. 17). Ed. Klaus Grubmüller and Markus Stock. Wiesbaden: Harrassowitz, 2003. 262.
496
Jörg Jochen Berns
the origin and center of the ceremonial, he must be able to distinguish the characteristics of the instrumentally produced sounds and operate in conformity to the rules, or intentionally break them in an innovative way.34 Even today, anyone socialized within the European cultural sphere succumbs to the suggestiveness of the sound symbolism that has been instilled in us over centuries, and which in baroque Europe was practiced and partly theorized in an almost encyclopedic way by the creation of the symphony orchestra. Thus, horns announce the hunt or the post, clarions call to battle, harps and flutes make water flow, violins call on the sky and play higher than all reason, while bassoon and oboe, saxophone and bagpipes transport us to faunic-bucolic spheres. This was still a function of film music, so that in the Hollywood of the 1930s the maxim “that the spectator should not be conscious of the music” could be formulated, against which Theodor W. Adorno and Hanns Eisler polemicized.35 The fact that the musical spheres created by such a sound theory were partly systematized in the baroque period or at least sorted in terms of topics, can be seen in the frontispiece of Kircher’s Phonurgia (fig. 5), which appeared in 1673. On the inscribed pedestal in the central semantic axis, is the female allegorical figure Phonurgia (the ‘soundmaker’). Above, with a double trumpet is the winged Fama, and further above, in a kind of throne of clouds, the haloed sign of the Holy Trinity. Gathered around this central axis are various spaces of resonance, identified by their own groups of musicians and particular musical instruments. At the lower left, a group of satyrs, herdsmen, and peasants blow on crumhorns and panpipes. Above, on a rocky hill, a lyre-playing Apollo is surrounded by muses playing flutes or shawm instruments and singing. At the lower right a solitary shepherd blows on a shawm, above is a hunt spurred on by the sound of horns, and above this a group of military cavalry leaps through a field accompanied by drums and trumpets. In the cove on the horizon, Neptune is surrounded by blaring tritons and naiads. Finally, along the upper edge of the print, inhabiting the cloudy 34
35
The prince is only sovereign in relation to particular aspects of the establishment and interruption of the ceremonial so that he can deliberately introduce irregularities to create an innovation in particular aspects of the ceremonial. Cf. Julius Bernhard von Rohr. Einleitung zur Ceremoniel-Wissenschafft Der Privat-Personen. Ed. Gotthardt Frühsorge, Leipzig, 1990 [facsimile of the edition Berlin, 1728]. 31ff., as well as Rohr. Einleitung zur Ceremoniel-Wissenschafft Der großen Herren. 10ff. Cf. Theodor W. Adorno and Hanns Eisler. Composing for the Films. Cambridge: Cambridge University Press, 1994. 9.
Instrumental Sound and Ruling Spaces of Resonance
497
Fig. 5: Music makers acting in various landscapes with various instruments. Frontispiece to Athanasius Kirchers Phonurgia nova (Kempten, 1673).
heavens spanning the whole print, is a heavenly orchestra already virtually in symphonic arrangement: left the strings, right the brass, and the organ in the center. What this frontispiece suggests is that every type of divertissement has its own acoustic profile ensured by particular musical instruments. On the earth – and particularly at court – various acoustic ambiences exist, various sound sceneries, for which each has an acoustic decorum, a certain sound that is appropriate and fitting for a particular place. In the heavens, however, such separations are not valid. Here, everything is symphonically and harmoniously united. Thus, the organ appears as a unifying element36 since, by virtue of its various voces and its mechani36
On the organ, cf. Hammerstein. Macht und Klang. 87ff., as well as Kircher. Neue Hall- vnd Thon-Kunst. 120ff.
498
Jörg Jochen Berns
cally adjustable tonal register, it alone achieves, without any other instruments, an optimum of sound production and different combinations of tonal color that could otherwise only be achieved by the addition of different instruments. Organ, Bell, Cannon In such a spectrum of instrumental sound production for ceremonial purposes, besides the quasi-universal sound synthesis of the organ, the bell and cannon represent the most powerful, because the furthest reaching but also most unmusical example of acoustic power demonstration. The cannon is the youngest in this group but also the most commonplace and brutal. It rivals both organ and bell. I want to make both rivalries clearer. a) On the Acoustic Rivalry between Organ and Cannon Organ and cannon relate to each other as complementary acoustic extremes. While the organ of the early modern period integrates by imitating all the tonal values of musical voices and instruments, and defines itself – at least phantasmatically – as sound reproducer per se, the cannonade eclipses every other noise. The organ makes every other instrument and every other musical intonation superfluous, while the cannonade makes every other acoustic articulation impossible. In terms of their sacral claim, organ and cannon also rival each other, though they complement each other as well.37 The cannon as “modern” long-range weapon (which after the fourteenth century led to a radical change in fortification systems) needed to be especially prepared and redesignated to be used as a sound instrument in ceremonial contexts at all. It was loaded in a special way to be able to fire shots – salutes – safely. The resulting dull, rolling, farechoing explosive roar was, in the early modern period, unanimously given the metaphor of ‘thunder.’ The cannonade was consequently interpreted as an atmospheric expression that was otherwise only avail37
“Even today the organ preserves, at a deeper level, so to speak, something of the old machine semantics; namely to be the depiction of heavenly cosmic harmonies, radiate numinous consecration, loftiness and power, or even to be the expression of the supra-personal per se.” Hammerstein. Macht und Klang. 98.
Instrumental Sound and Ruling Spaces of Resonance
499
able or attributable to God or the gods, generally to warn or intimidate mankind. (In this context it should be recalled that imitations of storms were already used in antiquity for the production and demonstration of sacral dignity.38 It should also be considered that in the early modern period the cannon competed with the organ in the imitation of storms. The baroque organ was occasionally provided with thunder machines, so that a storm could be made into an indoor event.39) In this sense the early modern powers, the “earthly gods” (Erdengötter),40 also used the cannonade as a sign of threat, by which the firing power of a fortification arrangement could be demonstrated and the strength of the defense as well as the prosperity of a commune, a fleet, a residence could be established. b) On the Acoustic Rivalry between Bell and Cannon The rivalry between bell and cannon was of a fundamentally different nature in that both aspired to a long-distance effect in public space while in the early modern period the organ was mostly employed as an indoor instrument and only rarely used outside as a transportable openair instrument.41 Bell ringing and cannonade as distantly echoing noises present today, in as much as they can still be heard at all, sounds from a distant world.42 Both once functioned as the territory-defining voices of a centralized power. Both shook the atmosphere, and still more the people, in a way that can no longer be properly conveyed to us. (Both were also occasionally produced by the same foundries.) 38 39
40 41
42
Ibid. 111ff. “In the echo organ of the court chapel of Lucerne there was a rain machine where a wooden drum would slowly rotate so that the gravel inside would slide, and in falling create the acoustic impression of rain. For the imitation of thunder, there were similar devices. Even in the second half of the eighteenth century, for example, they still played an important role in ‘storm fantasies’ or the pastoral plays of J.H. Knecht or Abbé Vogler.” Ibid. 102. On the origin and valency of the term “Erdengötter,” cf. the catalog Erdengötter. Berns et al., eds. XV. Because, in the Middle Ages, the organ was used as a courtly ceremonial instrument, and particularly in relation to the laudes regiae, it had to be mobile and transportable: “The laudes regiae took place during the reception of emperors and kings into towns or cities, at coronations and other ceremonial occasions . . . This also explains the fact that organs were brought along by ambassadors among others.” Hammerstein. Macht und Klang. 96. On this, cf. the empirically rich, theoretically stimulating presentation by Corbin. Village Bells. Passim.
500
Jörg Jochen Berns
The potency of the sound of bells and cannons as voices of centralized power was not nevertheless – this should be emphasized – exhausted in the function of defining the territory. Inherent in such sound signs was a tendency towards expansion that does not stop at territorial boundaries but transgresses them. Just such a boundary transgression and tendency towards an expansion of territory – and this is the important difference between acoustic and visual ceremonial – could not, however, be achieved by the usual visual (theatrical) media of the representation of power. Under the conditions of early modern politics with its crowded stage of different ruling rights and the broad extension of the territorium non clausum, the employment of bell and cannon could have an effect ‘inwards,’ homogenizing and leveling rule, while it signaled ‘outwards’ an operation into foreign territorial sovereignty. It would be worth investigating whether besides the heraldic claim there was not also something like an acoustic claim.43 Of course, the signal potential of bells, from a sacral and political point of view, is older and richer in variation than that of the cannon. For centuries (until they were suppressed by radio in the 1920s) bells provided the generally binding code of a communication system that conveyed and ensured individual and social identity. They regulated the relations between people according to a now forgotten rhythm, even those between life and death. If the cannonade could only signal a state of emergency, then this was only the case with bell ringing when either all bells were rung at once or only a single warning bell.44 This would then announce that time had fallen out of joint.
43
44
I borrow these considerations on the transgressive power and acoustic claim of bells and cannons from a letter from Volker Bauer (HAB Wolfenbüttel), whom I warmly thank again for these important details. Also the many medieval examples put forward by Sabine Zak speak for an acoustic claim: “In the Middle Ages, the thought is still alive that the spirit of God expresses itself aloud.” Zak. Musik als “Ehr und Zier.” 10. “A great peal of bells was part of the expression of a claim, it simultaneously demonstrated the power to accomplish this claim, above all against enemies.” Ibid. 13. This is how things proceeded during the days, weeks, or months reserved in Frankfurt for the election and coronation festivals. During this period, the prince electors would be called by the warning bell to the town hall and to make their way collectively to the election chapel, while after the coronation, the pealing of all the bells of Frankfurt Cathedral signaled the end of the state of emergency and the sealing off of the city. Cf. Wanger. Kaiserwahl und Krönung. 95, 101, and 120.
Instrumental Sound and Ruling Spaces of Resonance
501
EXCURS: On the Prescriptive Precision of the Instrument Definition One should of course ask by what right the organ, bell, and cannon might be defined as musical instruments, even as instruments at all. The organ, one could say, is not an instrument but a partly automated machine because it integrates many and potentially all musical instruments, thus making them superfluous – it is a universal and universalizing sound machine, and it tends towards automation. Bell and cannon on the other hand are not musical instruments in as much as their tonal volumes and scales cannot (or not strongly) be differentiated, and their sound cannot (or only in a limited way) be combined with the sounds of string or brass instruments. They cannot be differentiated, however, because, unlike (musical) instruments, they are not operated close to the body, or are not ‘at hand.’ They are not playable in two respects: either alone or in concerted action. Conversely, this means that instruments relate to the body. They are operated by means of body movements, and extend the performance of the body’s limbs or the five senses. Unlike machines they are simple, while machines combine several instruments and/or tools. Instruments differ from tools (which are likewise simple and related to the body) in that they obtain nothing, in a narrower sense produce nothing, but show something, make it experienceable, indicate or eliminate. At least this is the case for those groups or families of instruments that in English are largely described as such. Besides musical instruments there are measuring instruments, observation instruments, viewing instruments, investigation instruments, operating instruments, torture instruments, and so on. It is known that the differences between terms such as tool, instrument, implement, machine, apparatus, device in everyday practice, but also in scientific practice, are imprecise. This lack of precision is itself productive, as is proved by the multi-layered nature of the theorems and examples offered in this volume. If, however, one considers that in the efforts to define the term instrument, the species of musical instrument has over centuries (according to information on the history of lexicon definitions alone, since the sixteenth century) continuously played the most important role, the heuristic value of an investigation into the instrumentality of musical instruments cannot be denied.
502
Jörg Jochen Berns
Conclusion From the diverse observations of my historically and sociologically limited field of examples, ten rules, which with further empirical study could certainly become more precise and probably also more numerous, can be derived: 1) The prince is the center of the acoustic sphere. He appears almost as the central source of sound within a concentrically flowing sound pulsation. 2) Even though the prince need not produce the sounds of which he is the center himself, and usually does not produce them himself, he is still their cause. 3) When the prince is stationary, a dome of sound is created over him and around him. When he moves, a sound cocoon moves with him. 4) The princely sound source has an expansive tendency. The sound vacuoles surrounding the prince want to extend to the horizon, incessantly reaching out into space, defining borders, though also transgressing them. 5) The ceremonial centralization of sound has the purpose of drowning out contingent sounds, everyday noises, and keeping them at a distance from the prince. 6) The sounds that radiate out from the prince are of various aesthetic qualities. In the vicinity of the prince, they are softly and finely instrumented and well composed for a narrow circle of addressees. For the further and furthest addressees, on the other hand, coarse sounds are issued by means of loud, rough devices. 7) The changeability, mobility, and elasticity of ceremonial spaces are dependent on the local positioning and mobility of the instrument as well as the elasticity (modulation potential) and the volume of the instrumentally produced sounds. 8) All acoustic signs in the ceremonial are intented to make available to the senses something cognitively invisible. 9) The acoustic signs of the ruling ceremonial appear (almost) continually in combination with signs that call on other senses. 10) In the ceremonial, the senses of taste, smell, and even touch do not play such an important role as those of hearing and sight because they cannot cover (great) distances. The senses of sight and hearing on the other hand are active in a space-capturing and space-creating way, and only they can be instrumentalized; they therefore also correspond (complement or oppose) with each other in the modification of ceremonial spaces.
Instrumental Sound and Ruling Spaces of Resonance
503
The inspection of courtly-ceremonial possibilities and the instrumental conditions of acoustic signs that I have attempted to isolate here has opened insights into a systematic model that took many centuries to run its course. The instrumental musical, and extra-musical sounds created by devices that were used by the early modern prince, between the poles noise and silence, for the assertion of his potestas in the establishment and maintenance of space, seem so well calculated and observant of tradition, because the model of their interaction for the explanation of acoustic construction would be of use at least heuristically for later epochs. Accordingly, in relation to the early modern model, one could discuss if and then how the technical perfecting and extension of the acoustic signal-instrumentarium in the twentieth and twenty-first centuries accompanies and accelerates possibilities for the maintenance and extension of political control. It might then be asked whether such perfecting and extension of acoustic signaling does not actually create certain forms of control and possibly hinder others. Since the nineteenth century, with technical improvements in microphony and macrophony, with the enhancement of the way noises are technically preserved, almost randomly reproduced, as well as the way they are electronically simulated, the acoustic regulation of space has found new, previously undreamed of possibilities. Technical innovations have created new codes and made accessible new spaces of resonance of the most varied quality that are still control spaces. Despite this fact, one cannot fail to see that since the second half of the twentieth century at the latest there has been an enormous increase in the broad diffusion of noise as well as the distribution of visual signs. What this means, and whether this might lead to a gradual destruction of our musical instruments as well as our ability to hear, can be as little discussed here as the even more radical question of whether we are facing a time when the instrument will be totally replaced by the machine. This seems, however, to be the case.
Translation: Benjamin Carter
504
Jörg Jochen Berns
WORKS CITED Adorno, Theodor W. and Hanns Eisler. Composing for the Films. Cambridge: Cambridge University Press, 1994. Anonymous. “Beschreibung des prächtigen Festins, so wegen des Westphälischen Friedens an dem Fürstl. Hofe zu Weymar gehalten worden, de Anno 1650.” Johann Christian Lünig. Theatrum ceremoniale historico-politicum, oder Historisch- und Politischer Schau-Platz aller Ceremonien. Leipzig, 1719/20. Vol. I, 828-34. Anonymous. “Nachricht von der solennen Crönung König Georgens des I. in Groß-Britannien, de Anno 1714.” Johann Christian Lünig. Theatrum ceremoniale historicopoliticum, oder Historisch- und Politischer Schau-Platz aller Ceremonien. Leipzig, 1719/20. Vol. I, 1378-84. Bepler, Jill. “Ansichten eines Staatsbegräbnisses. Funeralwerke und Diarien als Quelle zeremonieller Praxis.” Zeremoniell als höfische Ästhetik in Spätmittelalter und Früher Neuzeit (= Frühe Neuzeit, vol. 25). Ed. Jörg Jochen Berns and Thomas Rahn. Tübingen: Niemeyer, 1995. 183-97. Berns, Jörg Jochen. “‘Vergleichung eines Vhrwercks, vnd eines frommen andächtigen Menschens.’ Zum Verhältnis von Mystik und Mechanik bei Spee.” Friedrich von Spee. Dichter, Theologe und Bekämpfer der Hexenprozesse. Ed. Italo Michele Battafarano. Gardolo di Trento: Reverdito, 1988. 101-206. Berns, Jörg Jochen. “Die Jagd auf die Nymphe Echo. Künstliche Echoeffekte in Poesie, Musik und Architektur der Frühen Neuzeit.” Die Mechanik in den Künsten. Studien zur ästhetischen Bedeutung von Naturwissenschaft und Technologie. Ed. idem and Hanno Möbius. Marburg: Jonas Verlag, 1990. 67-82. Berns, Jörg Jochen. “Kriegs- und Friedensbilder. Mittel ihrer ästhetischen Reflexion im 17. Jahrhundert.” Morgen-Glantz 9 (1999): 181-218. Berns, Jörg Jochen and Thomas Rahn. “Zeremoniell und Ästhetik.” Zeremoniell als höfische Ästhetik in Spätmittelalter und Früher Neuzeit (= Frühe Neuzeit, vol. 25). Ed. idem. Tübingen: Niemeyer, 1995. 650-66. Berns, Jörg Jochen et al., eds. Erdengötter. Fürst und Hofstaat in der Frühen Neuzeit im Spiegel von Marburger Bibliotheks- und Archivbeständen (= Schriften der UB Marburg, vol. 77). [Exhibit. cat.] Marburg: Universitätsbibliothek, 1997. Burckhardt, Martin. Metamorphosen von Raum und Zeit. Eine Geschichte der Wahrnehmung. Frankfurt a.M. and New York: Campus-Verlag, 1994. Bußmann, Klaus and Heinz Schilling, eds. 1648. Krieg und Frieden in Europa. 3 vols. Munich: Bruckmann, 1998. Canetti, Elias. Crowds and Power. Trans. Carol Stewart. New York: Continnum, 1973. Corbin, Alain. Village Bells. The Culture of the Senses in Nineteenth-Century French Countryside. Trans. Martin Thom. New York: Columbia University Press, 1998. Der Triumphzug Kaiser Maximilians I. 1516-1518. 147 Holzschnitte von Albrecht Altdorfer, Hans Burgkmair, Albrecht Dürer u.a. Mit dem von Kaiser Maximilian diktierten Programm (= Harenberg Kommunikation, vol. 100). Afterword by Horst Appuhn. Dortmund: Harenberg Kommunikation, 1979. Eckermann, Johann Peter. Gespräche mit Goethe in den letzten Jahren seines Lebens. Berlin: Aufbau-Verlag, 1982. Franke, Birgit. “Automaten in höfischen Lustgärten in der Frühen Neuzeit.” Automaten in Kunst und Literatur des Mittelalters und der Frühen Neuzeit (= Wolfenbütteler Mittelalter-Studien, vol. 17). Ed. Klaus Grubmüller and Markus Stock. Wiesbaden: Harrassowitz, 2003. 247-67. Haase, Rudolf. Johann Keplers Weltharmonik. Der Mensch im Geflecht von Musik, Mathematik und Astronomie. Munich: Diederichs, 1998.
Instrumental Sound and Ruling Spaces of Resonance
505
Hammerstein, Reinhold. Die Musik der Engel. Untersuchungen zur Musikanschauung des Mittelalters. Bern and Munich: Francke, 1962. Hammerstein, Reinhold. Macht und Klang. Tönende Automaten als Realität und Fiktion in der alten und mittelalterlichen Welt. Bern: Francke, 1986. Kircher, Athanasius. Musurgia universalis sive ars magna consoni et dissoni in X. libros digesta. 2 vols. Rome, 1650. [Kircher, Athanasius/Andreas Hirsch.] Artis magnae de consono & dißono ARS MINOR; Das ist/ Philosophischer Extract und Auszug/ aus deß Welt-berühmten Teutschen Jesuitens Athanasii Kircheri . . . MUSURGIA UNIVERSALI . . . Ausgezogen und verfertiget . . . von Andrea Hirschen/ . . . Evangel. Pfarrern zu Bächlingen . . . Leipzig: Zentralantiquariat der DDR, 1988 [facsimile of the edition Schwäbisch Hall, 1662]. [Kircher, Athanasius/Agathus Carion.] ATHANASII KIRCHERS è Soc: JESU Neue Hall- vnd Thon-Kunst/ Oder Mechanische Gehaim-Verbindung der Kunst und Natur . . . Jn unsere Teutsche Mutter-Sprach übersetzet von AGATHO CARIONE. Hanover: “Libri rari” Schäfer, 1984 [facsimile of the edition Nördlingen, 1684]. Kümmel, Werner. “Puls und Musik (16.-18. Jahrhundert).” Medizinhistorisches Journal 3 (1963): 269-93. Langgärtner, G. and G. May. “Akklamation.” Lexikon des Mittelalters. Munich: Deutscher Taschenbuch Verlag, 2002. Vol. I, 231f. Laufhütte, Harmut. “Das Friedensfest in Nürnberg 1650.” 1648. Krieg und Frieden in Europa. Ed. Klaus Bußmann and Heinz Schilling. Munich: Bruckmann, 1998. Vol. 2, 347-58. Lünig, Johann Christian. Theatrum ceremoniale historico-politicum, oder Historischund Politischer Schau-Platz aller Ceremonien. 2 [3] vols. Leipzig, 1719/20. Menestrier, Claude-François. Des décorations funèbres . . . Paris, 1683. Meyer, Rudolf J. Königs- und Kaiserbegräbnisse im Spätmittelalter. Von Rudolf von Habsburg bis zu Friedrich III. (= Forschungen zur Kaiser- und Papstgeschichte des Mittelalters, vol. 19). Cologne, Weimar, and Vienna: Böhlau, 2000. Quoika, Rudolf. Altösterreichische Hornwerke. Berlin: Merseburger, 1959. Rohr, Julius Bernhard von. Einleitung zur Ceremoniel-Wissenschafft Der Privat-Personen. Ed. Gotthardt Frühsorge. Leipzig, 1990 [facsimile of the edition Berlin, 1728]. Rohr, Julius Bernhard von. Einleitung zur Ceremoniel-Wissenschafft Der großen Herren. Ed. Monika Schlechte. Leipzig, 1990 [facsimile of the edition Berlin, 1733]. Seifert, Herbert. Die Oper am Wiener Kaiserhof im 17. Jahrhundert (= Wiener Veröffentlichungen zur Musikgeschichte, vol. 25). Tutzing: Schneider, 1985. Stieve, Gottfried. Europäisches Hoff-Ceremoniel. Leipzig: Gleditsch, 1715. Wanger, Bernd Herbert. Kaiserwahl und Krönung im Frankfurt des 17. Jahrhunderts. Darstellung anhand der zeitgenössischen Bild- und Schriftquellen und unter besonderer Berücksichtigung der Erhebung des Jahres 1612 (= Studien zur Frankfurter Geschichte, vol. 34). Frankfurt a.M.: Kramer, 1994. Wille, Günther. Musica romana. Die Bedeutung der Musik im Leben der Römer. Amsterdam: Schippers, 1967. Winterfeld, Friedrich Wilhelm von. Teutsche und Ceremonial-Politica. 2 vols. Frankfurt a.M. and Leipzig, 1700 and 1702. Wolfenbüttel Thesaurus for Early Modern Festival Books. Online source: http://www.hab.de/forschung/projekte/festkultur.htm. Zak, Sabine. Musik als “Ehr und Zier” im mittelalterlichen Reich. Studien zur Musik im höfischen Leben, Recht und Zeremoniell. Neuss: Päffgen, 1979.
About the Authors
BRUNO BACHIMONT Professor, Ph.D; engineer and philosopher. Teaching and research activities at the Université de Technologie de Compiègne; scientific director of the Institut National de l’Audiovisuel; member of the CNRS research unit “Heuristique et Diagnostic des systèmes complexes.” Main areas of study: epistemology (acquisition, modelling, and representation of knowledge); translation of epistemological results into computer systems; ontological structuring, indexing, and processing of meta-information; research at the intersection of artificial intelligence, terminology and linguistics, semiology, and documentation. Publications include: Le contrôle dans les systèmes à base de connaissances. Contribution à l’épistémologie de l’intelligence artificielle (Paris, 1992). JÖRG JOCHEN BERNS Professor emeritus of the Department of Modern German Literature and Media at the Philipps-Universität Marburg, Ph.D; director of the German Research Foundation (DFG)Project on mnemonics/ars memorativa. Main areas of study: late medieval and early modern cultural and intellectual history; German literature of the sixteenth to nineteenth century; court culture (court typology, courtly festivals, ceremony, theater); ars memorativa/ mnemonics; history of the newspaper and media; history of science; utopia research; academy history. Publications include: Die Herkunft des Automobils aus Himmelstrionfo und Höllenmaschine (Berlin, 1996); Erdengötter. Fürst und Hofstaat in der Frühen Neuzeit, im Spiegel von Marburger Bibliotheks- und Archivbeständen (co-ed., Marburg, 1997); Film vor dem Film. Bewegende und bewegliche Bilder als Mittel der Imaginationssteuerung in Mittelalter und Früher Neuzeit (Marburg, 2000); Seelenmaschinen. Gattungstraditionen, Funktionen und Leistungsgrenzen der Mnemotechniken vom späten Mittelalter bis zum Beginn der Moderne (co-ed., Vienna etc., 2000); Gedächtnislehren und Gedächtniskünste in Antike und Frühmittelalter (ed., Tübingen, 2003). OLAF BREIDBACH Professor, Ph.D, Sc.D; director of the Institute for the History of Medicine, Natural Science and Technology at the Friedrich-Schiller-Universität Jena; director of the ErnstHaeckel-Haus Museum; holds chair in history of natural sciences; in charge of “Theoretical Biology” at the biology-pharmaceutical faculty of the FSU Jena; member of the board of the research project “Ereignis Weimar-Jena – Kultur um 1800.” Main areas of study: theory of the history of sciences; system theory; scientific culture around 1800; natural philosophy; scientific perceptions. Publications include: Die Materialisierung des Ichs. Zur Geschichte der Hirnforschung im 19. und 20. Jahrhundert (Frankfurt a.M., 1997); Natur der Ästhetik, Ästhetik der Natur (Vienna, 1997); Das Anschauliche oder über die Anschauung von Welt. Ein Beitrag zur neuronalen Ästhetik (Vienna and New York, 2000); Naturwissenschaften um 1800 (co-ed., Weimar, 2001); Ästhetik – Hermeneutik –
508
About the Authors
Neurowissenschaften (co-ed., Münster, 2004); Bilder des Wissens. Zur Kulturgeschichte der wissenschaftlichen Wahrnehmung (co-ed., Munich, 2005); Naturphilosophie nach Schelling (co-ed., Stuttgart-Bad Cannstatt, 2005). MARTIN BURCKHARDT Ph.D; cultural and media theorist; teaching activities at various universities. Main areas of study: genealogy of appearance, politics of simulation. Radio pieces, essays, performances. Publications include: Metamorphosen von Raum und Zeit. Eine Geschichte der Wahrnehmung (Frankfurt a.M. and New York, 1994); Vom Geist der Maschine. Eine Geschichte kultureller Umbrüche (Frankfurt a.M. and New York, 1999); Brandlhuber. Eine Fiktion (Cologne, 2005); Die Scham der Philosophen (Berlin, 2006). CDs: Die Kehrseite der Vernunft (1987); Die Wüste oder das Elend der Avantgarde (1988); Change program please (1989); Die Offenbarung des Daniel Paul Schreber (1996); Das Seminar (1997). LORRAINE DASTON Professor, Ph.D; director at the Max Planck Institute for the History of Science, Berlin; honorary professor at the Humboldt-Universität Berlin. Main areas of study: history of objectivity, scientific attention, statistics, and probability. Publications include: The Probabilistic Revolution (co-ed., Cambridge, 1987); Classical Probability in the Enlightenment (Princeton, 1988); Wonders and the Order of Nature 1150-1750 (New York, 1998); Biographies of Scientific Objects (ed., Chicago, 2000); Eine kurze Geschichte der wissenschaftlichen Aufmerksamkeit (Munich, 2001); Wunder, Beweise und Tatsachen. Zur Geschichte der Rationalität (Frankfurt a.M., 2001); The Faces of Nature in Enlightenment Europe (co-ed., Berlin, 2003); Things that Talk. Object Lessons from Art and Science (ed., New York, 2004); The Moral Authority of Nature (co-ed., Chicago, 2004). GEORGES DIDI-HUBERMAN Ph.D; philosopher and art historian. Teaches at the École des Hautes Études en Sciences Sociales in Paris. Main areas of study: history and theory of the image; anthropology; psychoanalysis; space and time in art. Publications include: Fra Angelico. Dissemblance and Figuration (Chicago, 1995); Devant l’image. Question posée aux fins d’une histoire de l’art (Paris, 1990); Ce que nous voyons, ce qui nous regarde (Paris, 1992); Ouvrir Venus. Nudité, rêve, cruauté (ed., Paris, 1999); L’image survivante. Histoire de l’art et temps des fantômes selon Aby Warburg (Paris, 2002); Ninfa moderna. Essai sur le drapé tombé (ed., Paris, 2002); Invention of Hysteria. Charcot and the Photographic Iconography of the Salpêtrière (Cambridge, MA, 2004); Images malgré tout (Paris, 2003); Confronting Images. Questioning the Ends of a Certain History of Art (Philadelphia, 2005) STEFAN DITZEN Ph.D; German Research Foundation-scholar (DFG) of the research training group “Bild, Körper, Medium. Eine anthropologische Perspektive” at the Hochschule für Gestaltung, Karlsruhe. Main areas of study: technical and socio-psychological conditions for microscopic visualization – stages of a pictorial history of the microscope; perceptive pattern of microscopic visualization. Publications include: “Mikrofotografien mit Hilfe des Raster-Elektronenmikroskops. Neue Wunderkammern im Mikrokosmos.” Palast des Wissens. Ed. Brigitte Buberl and Michael Dückershoff (2 vols. Munich, 2003). FRANK FEHRENBACH Professor of History of Art and Architecture at Harvard University, Ph.D. Main areas of study: liveliness as an aesthetic category in the visual arts of the thirteenth-eighteenth century; Leonardo da Vinci’s studies on movement, perception, and mechanics; natural
About the Authors
509
philosophy of the Renaissance; iconology of the early modern fountain. Publications include: Die Goldene Madonna im Essener Münster. Der Körper der Königin (Ostfildern, 1996); Licht und Wasser. Zur Dynamik naturphilosophischer Leitbilder im Werk Leonardo da Vincis (Tübingen, 1997); Leonardo da Vinci. Natur im Übergang. Beiträge zu Wissenschaft, Kunst und Technik (ed., Munich, 2002); Compendia mundi. Gianlorenzo Berninis Fontana dei Quattro Fiumi (1648-51) und Nicola Salvis Fontana di Trevi (1732-62) (Munich and Berlin, 2007). PETER GALISON Professor of History of Science and Physics at Harvard University, Ph.D. Main areas of study: philosophy and history of modern physics, focussing on experiment, instrument, and theory; theory of relativity at the intersection of technology, philosophy, and physics; architecture of science. Publications include: How Experiments End (Chicago, 1987); Big Science. The Growth of Large-Scale Research (ed., Stanford, 1992); Image and Logic. A Material Culture of Microphysics (Chicago and London, 1997); Picturing Science, Producing Art (co-ed., New York, 1998); The Architecture of Science (co-ed., Cambridge, 1999); Atmospheric Flight in the 20th Century (co-ed., Dordrecht, 2000); Einsteins Uhren, Poincarés Karten (Frankfurt a.M., 2003); Scientific Authorship (co-ed., New York, 2003). THOMAS F. GIERYN Professor of Sociology and senior lecturer in Philosophy and History of Science at Indiana University, Bloomington, Ph.D. Main areas of study: sociology of science; research-field selection, and authority claim in science; topography of knowledge. Publications include: Theories of Science in Society. Science, Technology, and Society (co-ed., Bloomington, 1990); Science and Social Structure. A Festschrift for Robert K. Merton (co-ed., New York, 1993); Cultural Boundaries of Science. Credibility on the Line (Chicago, 1999). GERALD HARTUNG Associate professor, Ph.D; research assistent at the Forschungsstätte der Evangelischen Studiengemeinschaft e.V. (Heidelberg); research fellow at the Max-Weber-Kolleg, University of Erfurt and lecturer at the University of Heidelberg. Main areas of study: political philosophy, philosophical anthropology / philosophy of culture, theory of intellectual and cultural studies. Publications include: Zwischen Narretei und Weisheit. Biographische Skizzen und Konturen alter Gelehrsamkeit (co-ed., Hildesheim and New York, 1997); Ernst Cassirer. Die Philosophie der Aufklärung (ed., Hamburg, 1998); Die Naturrechtsdebatte. Geschichte der Obligatio vom 17. bis 20. Jahrhundert (Freiburg, 1998); Das Maß des Menschen. Aporien der philosophischen Anthropologie und ihre Auflösung in der Kulturphilosophie Ernst Cassirers (Weilerswist, 2003); Weltoffener Humanismus. Lebens- und Ideengeschichten deutsch-jüdischer Humanisten des 20. Jahrhunderts (Bielefeld, 2006). JOCHEN HENNIG M.Sc (physics). Research associate at the Hermann von Helmholtz-Zentrum für Kulturtechnik Berlin, project “Das Technische Bild.” Main areas of study: experiment and image; visual communication of knowledge; science exhibitions. Publications include: Der Spektralapparat Kirchhoffs und Bunsens (Berlin and Munich, 2003); “Changes in the Design of Scanning Tunneling Microscopic Images from 1980 to 1990.” Techné 8.2 (2004); “Vom Experiment zur Utopie. Bilder in der Nanotechnologie.” Instrumente des Sehens (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 2.2). Ed. Horst Bredekamp and Gabriele Werner (Berlin, 2004).
510
About the Authors
DON IHDE Professor of Philosophy at State University of New York, Stony Brook, Ph.D. Main areas of study: philosophy of technology; philosophy of the natural sciences; intercultural perception; representational technologies; feminist critique of science and technology; social and political dimensions of the sciences; body theories. Publications include: Technology and the Lifeworld. From Garden to Earth (Indianapolis, 1990); Instrumental Realism. The Interface between Philosophy of Science and Philosophy of Technology (Bloomington, 1991); Philosophy of Technology. An Introduction (New York, 1993); Expanding Hermeneutics. Visualism in Science (Evanston, 1998); Bodies in Technology (Minneapolis, 2002). SYBILLE KRÄMER Professor at the Institute for Philosophy of the Freie Universität Berlin, Ph.D; Permanent Fellow at the Wissenschaftskolleg Berlin. Main areas of study: philosophy and mathematics in the seventeenth century; philosophy of language; media theory; basic questions of cultural theory. Publications include: Symbolische Maschinen. Die Idee der Formalisierung in geschichtlichem Abriß (Darmstadt, 1988); Berechenbare Vernunft. Kalkül und Rationalismus im 17. Jahrhundert (Berlin and New York, 1991); Sprache – Sprechakt – Kommunikation. Sprachtheoretische Positionen im 20. Jahrhundert (Frankfurt a.M., 2001); Bild – Schrift – Zahl (co-ed., Munich, 2003); Performativität und Medialität (ed., Munich, 2004); Schrift. Kulturtechnik zwischen Auge, Hand und Maschine (co-ed., Munich, 2005); Verletztende Worte. Die Grammatik sprachlicher Missachtung (co-ed., Bielefeld, 2007); Stimme (co-ed., Frankfurt a.M., 2006); Medium, Bote, Übertragung. Kleine Metaphysik der Medialität (forthcoming). JAN LAZARDZIG Ph.D. Research associate at the Institute for Theater Studies of the Freie Universität Berlin. Main area of study: Baroque theater; theatricality in the history of culture and science of the seventeenth century. Publications include: Kunstkammer, Laboratorium, Bühne. Schauplätze des Wissens im 17. Jahrhundert (co-ed., Berlin and New York, 2003); English version: Collection, Laboratory, Theater. Scenes of Knowledge in the 17th Century (Berlin and New York, 2005); Instrumente in Kunst und Wissenschaft. Zur Architektonik kultureller Grenzen im 17. Jahrhundert (co-ed., Berlin and New York, 2006); Spektakuläre Experimente. Praktiken der Evidenzproduktion im 17. Jahrhundert (co-ed., Berlin and New York, 2006); Theatermaschine und Festungsbau. Paradoxien der Wissensproduktion im 17. Jahrhundert (Berlin, 2007). ANGELA MAYER-DEUTSCH MA. Research associate at the Hermann von Helmholtz-Zentrum für Kulturtechnik Berlin, project “Das Technische Bild.” Main area of study: image and knowledge around 1700. Publications include: “Iconographia Kircheriana.” Athanasius Kircher. Il Museo del Mondo. Ed. Eugenio Lo Sardo (Rome, 2001); “Quasi-optical Palingenesis. The Circulation of Portraits and the Image of Kircher.” Athanasius Kircher, S.J. The Last Man who Knew Everything. Ed. Paula Findlen (London and New York, 2004). DIETER MERSCH Professor of Media Studies at the Universität Potsdam, Ph.D. Main areas of study: media philosophy; philosophy of art; semiotics; philosophy of language; aesthetics; philosophy of the nineteenth and twentieth centuries. Publications include: Welten im Kopf. Profile der Gegenwartsphilosophie (co-ed., Hamburg, 1996); Zeichen über Zeichen (Munich, 1998); Ereignis und Aura. Untersuchungen zu einer performativen Ästhetik (Frankfurt
About the Authors
511
a.M., 2001); Was sich zeigt. Materialität, Präsenz, Ereignis (Munich, 2002); Performativität und Praxis (ed., Munich, 2003). ANDREAS MEYER Associate professor, Ph.D; lecturer in the Department of Musicology and Music Education at the Johann Wolfgang Goethe-Universität, Frankfurt am Main; member of the “Study Group on Folk Musical Instruments” at the International Council for Traditional Music. Main areas of study: ethno-organology; music in the museum; African and AfricanAmerican music; local pop music. Publications include: Der traditionelle Calypso auf Trinidad (Hamburg, 1991); Afrikanische Trommeln. West- und Zentralafrika (Berlin, 1997); Überlieferung, Individualität und musikalische Interaktion. Neuere Formen der Ensemblemusik in Asante/Ghana (Frankfurt a.M., 2005). FLORIAN NELLE Associate professor, Ph.D; part-time lecturer at the Institute of Theater Studies of the Freie Universität Berlin. Main areas of study: aesthetic dimensions of science in the seventeenth century, particularly the relationship between mannerist poetry and experimental science; history of artificial paradises (world-theater, landscape garden, world exhibition, movie palace). Publications include: Exzentrische Räume (co-ed., Stuttgart, 2000); Bühnen Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst (co-ed., Berlin, 2003); Künstliche Paradiese (Berlin, 2005). CONNY RESTLE Professor, Ph.D; director of the Musikinstrumenten-Museum SIMPK Berlin; lecturer at the Freie Universität Berlin. Main areas of study: musical instruments of antiquity and the Middle Ages; musical instruments of the sixteenth, seventeenth, and eighteenth centuries. Publications include: Bartolomeo Cristofori und die Anfänge des Hammerclaviers (Munich, 1989); Faszination Klavier. 300 Jahre Pianofortebau in Deutschland (Munich, 2002); Richard Strauß im kaiserlichen Berlin (Berlin, 2001); In aller Munde. Mundharmonika, Handharmonika, Harmonium. Eine 200jährige Erfolgsgeschichte (Berlin, 2002). HANS-JÖRG RHEINBERGER Sc.D; director at the Max Planck Institute for the History of Science, Berlin; honorary professor in the Institute of Philosophy, Theory of Science, History of Science and Technology at the Technische Universität Berlin. Main areas of study: history of the natural sciences; experimental systems und spaces of knowledge. Publications include: Experimentalisierung des Lebens. Experimentalsysteme in den biologischen Wissenschaften 1850/1950 (ed., Berlin, 1993); Räume des Wissens. Repräsentation, Codierung, Spur (ed., Berlin, 1997); Experimentalsysteme und epistemische Dinge (Göttingen, 2001); Reworking the Bench. Research Notebooks in the History of Science (ed., Dordrecht, 2003); Classical Genetic Research and its Legacy. The Mapping Cultures of Twentieth-Century Genetics (ed., London, 2004); Iterationen (Berlin, 2005); Epistemologie des Konkreten (Frankfurt a.M., 2006). HELMAR SCHRAMM Professor at the Institute for Theater Studies of the Freie Universität Berlin, Ph.D. Main area of study: theater culture in the field of tension between history of media and science. Publications include: Karneval des Denkens. Theatralität im Spiegel philosophischer Texte des 16. und 17. Jahrhunderts (Berlin, 1996); Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst (ed., Berlin, 2003); Kunstkammer, Laboratorium,
512
About the Authors
Bühne. Schauplätze des Wissens im 17. Jahrhundert (co-ed., Berlin and New York, 2003); English Version: Collection, Laboratory, Theater. Scenes of Knowledge in the 17th Century (Berlin and New York, 2005); Instrumente in Kunst und Wissenschaft. Zur Architektonik kultureller Grenzen im 17. Jahrhundert (co-ed., Berlin and New York, 2006); Spektakuläre Experimente. Praktiken der Evidenzproduktion im 17. Jahrhundert (co-ed., Berlin and New York, 2006). LUDGER SCHWARTE Assistant professor for “Theory of Images”, Ph.D. Teaches philosophy at the University of Basel. Main area of study: philosophy of science architecture. Publications include: Die Regeln der Intuition. Kunstphilosophie nach Adorno, Heidegger und Wittgenstein (Munich, 2000); Kunst als Strafe. Zur Ästhetik der Disziplinierung (co-ed., Munich, 2003); Körper und Recht. Anthropologische Dimensionen der Rechtsphilosophie (co-ed., Munich, 2003); Kunstkammer, Laboratorium, Bühne. Schauplätze des Wissens im 17. Jahrhundert (co-ed., Berlin and New York, 2003); English version: Collection, Laboratory, Theater. Scenes of Knowledge in the 17th Century (Berlin and New York, 2005); Instrumente in Kunst und Wissenschaft. Zur Architektonik kultureller Grenzen im 17. Jahrhundert (co-ed., Berlin and New York, 2006); Spektakuläre Experimente. Praktiken der Evidenzproduktion im 17. Jahrhundert (co-ed., Berlin and New York, 2006); H. OTTO SIBUM Hans Rausing Professor of History of Science, director of the “Office for History of Science” at Uppsala University. Main areas of study: history of experimental knowledge from seventeenth to early twentieth century; non-written records in the physical sciences: objects, practice-bound knowledge, experience, body technologies, practices of theorization. Publications include: Das fünfte Element. Wirkungen und Deutungen der Elektrizität (co-author, Reinbeck, 1987); Physik aus ihrer Geschichte verstehen. Entstehung und Entwicklung naturwissenschaftlicher Denk- und Arbeitsstile in der Elektrizitätsforschung des 18. Jahrhunderts (Wiesbaden, 1990); “Les gestes de la mesure. Joule, les pratiques de la brasserie et la science.” Annales: histoire, sciences sociales 53 (4-5, 1998); “Il numero d’oro del secolo. Storia di un fatto scientifico.” Quaderni storici 108 (3, 2001); Instruments, Travel and Science. Itineraries of Precision from the Seventeenth to the Twentieth Century (co-ed., London, 2002); Scientific Personae (co-ed., Cambridge, 2003); “Wissen aus erster Hand. Mikrodynamik wissenschaftlichen Wandels im frühviktorianischen England.” Historische Anthropologie 13.3 (2005); The Heavens on Earth. Observatory Techniques in the Nineteenth Century (co-ed, forthcoming Durham, 2008). BARBARA MARIA STAFFORD Professor in the Department of Art History at the University of Chicago, Ph.D. Main areas of study: art and image theories from the sixteenth century to the romantic period; modern media and the interface between art and science in the modern period. Publications include: Symbol and Myth. Humbert de Superville’s Essay on Absolute Signs in Art (Cranbury, 1979); Voyage into Substance. Art, Science, Nature and the Illustrated Travel Account, 1790-1840 (Cambridge, 1984); Artful Science. Enlightenment Entertainment and the Eclipse of Visual Education (Cambridge, 1994); Good Looking. Essays on the Virtue of Images (Cambridge, 1996); Devices of Wonder. From the World in a Box to Images on a Screen (exhibit.cat. Getty Museum, Los Angeles, 2001); Beyond Productivity. Information Technology, Innovation, and Creativity (Washington, 2003); Echo Objects. The Cognitive Work of Images (Chicago, 2007).
About the Authors
513
NICOLA SUTHOR Ph.D: research associate at the Kunsthistorische Institut in Florenz (Max Planck Institut). Main areas of study: virtuosity and capriciousness in the painting of the pre-modern period; terms of mastery in the modern period. Publications include: “Portät.” Geschichte der klassischen Bildgattungen in Quellentexten und Kommentaren (co-ed., Berlin, 1999); Augenlust bei Tizian. Zur Konzeption sensueller Malerei in der Frühen Neuzeit (Munich, 2003); Ansteckung. Zur Körperlichkeit eines ästhetischen Prinzips (co-ed., Munich, 2005); Verklärte Körper. Ästhetik der Transfiguration (co-ed., Munich, 2006). GERHARD WIESENFELDT Ph.D, M.Sc (physics); lecturer in the Programme for History and Philosophy of Science at the University of Melbourne. Main areas of study: conceptions of Romanticism in the natural sciences; experimental sciences from the seventeenth to the early nineteenth century; sciences in early-modern universities; visual culture of the natural sciences. Publications include: Wahrnehmung der Natur – Natur der Wahrnehmung. Studien zur Geschichte visueller Kultur um 1800 (co-ed., Amsterdam and Dresden, 2001); Leerer Raum in Minervas Haus. Experimentelle Naturlehre an der Universität Leiden, 1675-1715 (Amsterdam, 2002).
Image Credits Rheinberger: (Fig. 1) O. Langendorff. Physiologische Graphik. Ein Leitfaden der in der Physiologie gebräuchlichen Registrirmethoden. Leipzig and Vienna: Deuticke, 1891; (Fig. 2) T.H. Morgan, C.B. Bridges, and A.H. Sturtevant. “The Genetics of Drosophila.” Biobliographia Genetica. Vol. II. Ed. J.P. Lotsy and H.N. Kooiman. ’S-Gravenhage: Nijhoff, 1925; (Fig. 3) A. Yonath, W. Bennett, S. Weinstein and H.G. Wittmann. “Crystallography and Image Reconstructions of Ribosomes.” The Ribosome. Structure, Function & Evolution. Ed. W.E. Hill et al. Washington: Amer Society, 1990; (Fig. 4) G. Stöffler et al. “Structural Organization of the Escherichia coli Ribosome and Localization of Functional Domains.” Ribosomes. Structure, Function, and Genetics. Ed. G. Chambliss et al. Baltimore: University Park Press, 1980; (Fig. 5) With the kind permission of Edward F. Polic. Breidbach: (Fig. 1 and 4) Robert Fludd. Utriusque Cosmi Maioris scilicet et MINORIS METAPHYSICA, PHYSICA ATQUE TECHNICA HISTORIA. Oppenheim, 1617; (Fig. 2, 3 and 5-9) Robert Fludd. Microcosmi historia . . . Oppenheim, 1619. Fehrenbach: (Fig. 1) Leonardo da Vinci. Deluge of Tools. Pen and black chalk, ca. 1510-15. Windsor, Royal Library, No. 12698 recto; (Fig. 2) Peter Apian. Practica for 1532. Landshut 1531; (Fig. 3) Jacopo Pontormo. Alessandro de‘ Medici. Oil on wood panel, 1534. Philadelphia Museum of Art, Johnson Collection; (Fig. 4) Leonardo da Vinci. Sickle Car. Pen and inc, ca. 1490. Turin, Biblioteca Reale, No. 15583; (Fig. 5) Leonardo da Vinci. Military Project. Pen and ink, ca. 1503/04. Windsor, Royal Library, No. 12275; (Fig. 6) Agostino Ramelli. Diverse et artificiose machine. Paris, 1588 (fig. CXLVI); (Fig. 7) Leonardo da Vinci. Cannon Foundry. Pen and ink, ca. 1495. Windsor, Royal Library, No. 12647; (Fig. 8) Leonardo da Vinci. Project of a Dredger. Pen and ink, ca. 1500. Codex Atlanticus, fol. 4 recto. Milan, Biblioteca Ambrosiana; (Fig. 9) Leonardo da Vinci. Map of the Arno. Various media, 1503/04. Windsor, Royal Library, No. 12279; (Fig. 10) Leonardo da Vinci. Study for the “Battle of Anghiari.” Black chalk, ca. 1504. Budapest, Szépmüvészeti Múzeum, No. 1774; (Fig. 11) Leonardo da Vinci. Study for the „Battle of Anghiari.” Red chalk, ca. 1504. Budapest, Szépmüvészeti Múzeum, No. 1775; (Fig. 12) Leonardo da Vinci. Study of a Head. Silver point, ca. 1485. Turin, Biblioteca Reale, No. 15572; (Fig. 13) Leonardo da Vinci. Spring. Pen and ink, ca. 1492. Codex Madrid I, fol. 45r (Madrid, Biblioteca Nacional). Suthor: (Fig. 1) Nicolas Régnier. Self-Portrait with the Portrait of Vincenzo Giustiniani on the Easel. Oil on canvas, 1623-1624. Cambridge, Fogg Art Museum (Harvard University Art Museum); (Fig. 2) Simon Vouet. Self-Portrait. Oil on canvas, ca. 1620. Arles, Musée Réattu; (Fig. 3) Simon Vouet. „Le Spadassin“. Oil on canvas, ca. 1620. Brunswick, Herzog Anton Ulrich Museum. Lazardzig: (Fig. 1) Heinrich Zeising. Theatri machinarum [6 parts]. 3 vols. Leipzig, 1607-1614, part II, fig. 12; part IV, fig. 6. With the kind permission of the Herzog August Bibliothek, Wolfenbüttel (sign. N 80a. 4o); (Fig. 2) Agostino Ramelli. Schatzkammer/ Mechanischer Künste/ . . . Leipzig, 1620. 103, 179. With the kind permission of the Herzog August Bibliothek, Wolfenbüttel (sign. M: Od 4º 48); (Fig. 3) Salomon de Caus. Les raisons des forces mouvantes . . . Frankfurt a.M., 1615, fol. 35r. With the kind permission of the Herzog August Bibliothek, Wolfenbüttel
516
Image Credits
(sign. A: 5 Geom. 2o (2)). Mayer-Deutsch: (Fig.1) Francisco de Florencia. Historia de la Provincia de la Compania de Jesus da Nueva-Espania. Mexiko, 1694. With the kind permission of the Staatsbibliothek zu Berlin; (Fig. 2-5) Athanasius Kircher. Ars Magna Lucis et Umbrae. Amsterdam 1671. 130, 783, 770, 718. With the kind permission of the Staatsbibliothek zu Berlin. Restle: (Fig. 1 and 2) Michael Praetorius. Syntagma Musicum. De Organographia. Wolfenbüttel, 1620 (plate supplement). With the kind permission of the Herzog August Bibliothek, Wolfenbüttel, (sign. 1.1 Musica (2)); (Fig. 3-5) Athanasius Kircher: Musurgia universalis sive ars magna consoni et dissoni in X libros digesta. Rome, 1650. With the kind permission of the Herzog August Bibliothek, Wolfenbüttel, (sign. Musica fol. 1.2). Meyer: (Fig. 1) © Gerd Grupe; (Fig. 2 and 3) Michael Praetorius. Syntagma musicum. De Organographia. Wolfenbüttel, 1620 (plate supplement), plate XXXI and XXX. With the kind permission of the Herzog August Bibliothek, Wolfenbüttel; (Fig. 4) © Andreas Meyer. Sibum: (Fig. 1) Whipple Museum for the History of Science, Cambridge University, UK; (Fig. 2) Torpedo fish: John Hunter. “Anatomical Observations on the Torpedo.” Philosophical Transactions 63 (1773-74): 488; Electrical column by Volta: © Musée des arts et métiers – CNAM, Paris; (Fig. 3) L. Figuier. Les merveilles de la science ou description populaire des inventions modernes. Paris, 1868. Vol. 2, 389, fig. 238. Ditzen: (Fig. 1) Mara Miniati, ed. Museo di Storia della Scienza. Catalogo. [Exhibit. cat.] Florenz: Olschki, 1991; (Fig. 2) Christian Gottlieb Hertel: Vollständige Anweisung zum Glass-Schleiffen . . . Halle, 1716; (Fig. 3 left; Fig. 4 upper left; Fig. 5 left) Robert Hooke. Micrographia . . . New York, 1961 [facsimile of the edition London, 1665]; (Fig. 3 right) Johann Frantz Griendel. Micrographia nova . . . Nuremberg, 1687; (Fig. 4 upper right) Philippo Bonanni. Observationes circa viventia . . . Cum Micrographia curiosa. Rome, 1699; (Fig. 4 lower left) Louis Joblot. Observations d’histoire naturelle. Paris, 1754; (Fig. 4 lower right) Johannes Jacob Scheuchzer. Physica sacra. Augsburg and Ulm, 1731-1735; (Fig. 5 right) Gerard L’E. Turner. Essays on the History of the Microscope. Oxford: Senecio Publications, 1980. Hennig: (Fig. 1) With kind permission from Physical Review Letters 50,2 (1982): 120-123, fig. 1. © American Physical Society; (Fig. 2) With kind permission from Europhysics Letters 1,1 (1986): 3136; (Fig. 3a, b) With kind permission from Physical Review Letters 52,15 (1984): 13041307, Fig. 1,2. © American Physical Society; (Fig. 4 a,b) © Hartwig Thomas; (Fig. 5) Erich Stoll. “Information- and Image-Processing of Scanning Tunneling Microscope Data?” SPIE 599 (1985): 442-450; (Fig. 6) With kind permission from Nature 344 (1990): 524-6. © Macmillan Publishers Ltd. Ihde: (Fig. 1) Germanisches Nationalmuseum Nürnberg, ed. Albrecht Dürer. Das druckgraphische Werk. Munich: Prestel, 2004. 265; (Fig. 2) Athanasius Kircher. Ars magna lucis et umbrae. Rome, 1646; (Fig. 3 and 4) © Don Ihde. Didi-Huberman: (Fig. 1-4) Michel Frizot. Nouvelle histoire de la photographie. Paris et al., 1995. 152, 78, 95, 156. Berns: (Fig. 1) Athanasius Kircher Musurgia Universalis. 2 vols. Rome, 1650. With the kind permission of Herzog August Bibliothek, Wolfenbüttel (sign. Musica fol. 1.2); (Fig. 2-5) Athanasius Kircher. Itinerario des Éxtasis o las Imágenes de un saber Universal. Ed. Gómez de LinaĖo. 2 vols. Madrid, 1986. 122-130.
Bibliography Abbott, Andrew. Department and Discipline. Chicago Sociology at One Hundred. Chicago: University of Chicago Press, 1999. Académie des Sciences, Paris. Passage de Vénus du décembre 1882. Rapports préliminaires. Paris: Gauthier-Villars, 1883. Adami, Giuseppe. “L’ingegnere-scenografo e l’ingegnere-venturiero.” Barocke Inszenierung. Ed. Rudolf Preimesberger and Joseph Imorde. Emsdetten and Zurich: Edition Imorde, 1999. 159-85. Adolphe, Lydie. La dialectique des images chez Bergson. Paris: P.U.F, 1951. Adorno, Theodor W. Kant’s Critique of Pure Reason. Ed. Rolf Tiedemann. Trans. Rodney Livingstone. Cambridge: Polity Press, 2001. Adorno, Theodor W. and Hanns Eisler. Composing for the Films. Cambridge: Cambridge University Press, 1994. Agricola, Georg. Vom Bergkwerck 12 Bücher . . . Basel, 1557. Ahrens, Wilhelm, ed. Briefwechsel zwischen Carl Gustav Jacob Jacobi und Moritz Hermann Jacobi. Leipzig: Teubner, 1907. Airy, George Biddell, ed. Account of Observations of the Transit of Venus, 1874, December 8, Made under the Authority of the British Government. And of the Reduction of the Observations. London: Her Majesty’s Stationery Office, 1881. Alberti, Leon Battista. On Painting. Trans. Cecil Grayson. Ed. Martin Kemp: London: Penguin, 1991. Algarotti, Francesco. Saggio sopra la Pittura. Ed. William Spaggiari. Rome: Archivio Guido Izzi, 2000. Alkmaion. “24, A 10.” Die Fragmente der Vorsokratiker. Ed. Hermann Diels and Walther Kranz. Berlin: Weidmann, 1951. Allen, Thomas J. Managing the Flow of Technology. Cambridge: MIT Press, 1977. Alloa, Emanuel. Diaphanes. Vom Begriff zum Phänomen der ikonischen Differenz. Master’s thesis at the Institut für Philosophie of the Freie Universität, Berlin, 2003. Ames-Lewis, Francis. La matita nera nella pratica di disegno di Leonardo da Vinci (= Lettura Vinciana, vol. 41). Florence: Giunti, 2002. Anderson, Robert G.W. “The Archaeology of Chemistry.” Instruments and Experimentation in the History of Chemistry. Ed. Frederic L. Holmes and Trevor H. Levere. Cambridge: MIT Press, 2000. Andréadès, Andreas Michael. History of the Bank of England 1640-1903. London: P.S. King, 1966. Anonymous. “Beschreibung des prächtigen Festins, so wegen des Westphälischen Friedens an dem Fürstl. Hofe zu Weymar gehalten worden, de Anno 1650.” Johann Christian Lünig. Theatrum ceremoniale historico-politicum, oder Historisch- und Politischer Schau-Platz aller Ceremonien. Leipzig, 1719/20. Vol. I, 828-34. Anonymous. “Nachricht von der solennen Crönung König Georgens des I. in Groß-Bri-
518
Bibliography
tannien, de Anno 1714.” Johann Christian Lünig. Theatrum ceremoniale historicopoliticum, oder Historisch- und Politischer Schau-Platz aller Ceremonien. Leipzig, 1719/20. Vol. I, 1378-84. Anonymous. “The Ballad of Gresham College.” Ed. Dorothy Stimson. Isis 18 (1932): 103-17. Apuleius. The Apologia and Florida of Apuleius of Madaura. Trans. H.E. Butler. Westport: Greenwood Press, 1970. Apuleius. The Golden Ass or Metamorphoses. Trans. E.J. Kenney. London: Penguin Books, 1998. Aristotle. De partibus animalium/Parts of animals. Trans. A.L. Peck. Cambridge: Cambridge University Press, 1983. Ashton, Dore. The New York School. A Cultural Reckoning. New York: Penguin, 1972. Ashton, Dore. “Introduction.” William C. Seitz. Abstract Expressionist Painting in America. Cambridge: Harvard University Press, 1983. XV-XXVI. Ashworth, William B., Jr. “Natural History and the Emblematic World View.” Reappraisals of the Scientific Revolution. David C. Lindberg and Robert S. Westman. Cambridge: Cambridge University Press, 1990. 303-32. Aubignac, François Hédelin d’. La pratique du théâtre. Amsterdam, 1715. Aubignac, François Hédelin d’. La pratique du théâtre. Munich: Fink, 1971. Augustine. The City of God against the Pagans. 7 vols. Trans. David S. Wiesen. Cambridge: Harvard University Press, 1968. Augustinus, Aurelius. De musica. Ed. and trans. Frank Hentschel. Hamburg: Meiner, 2002. Austin, John L. How to Do Things with Words. Cambridge: Harvard University Press, 2005. Auwers, Arthur, ed. Die Venus-Durchgänge 1874 und 1882. Bericht über die Deutschen Beobachtungen. 6 vols. Berlin, 1887-98. Bach, Thomas. “‘Was ist das Thierreich anders als der anatomirte Mensch …?’ Oken in Göttingen (1805-1807).” Lorenz Oken (1779-1851). Ein politischer Naturphilosoph. Ed. Olaf Breidbach, Hans-Joachim Fliedner, and Klaus Ries. Weimar: Böhlau, 2001. 73-91. Bachelard, Gaston. Le rationalisme appliqué. Paris: P.U.F., 1998. Bacher, Jutta. “Artes Mechanicae.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 35-49. Bacher, Jutta. “‘Ingenium vires superat.’ Die Emanzipation der Mechanik und ihr Verhältnis zu Ars, Scientia und Philosophia.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 519-55. Bacher, Jutta. “Das Theatrum machinarum. Eine Schaubühne zwischen Nutzen und Vergnügen.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. Hans Holländer. Berlin: Mann, 2000. 509-18. Bachimont, Bruno. Herméneutique matérielle et artéfacture. Des machines qui pensent aux machines qui donnent à penser. Ph.D. École Polytechnique, Paris, 1996. Bachimont, Bruno. “L’intelligence artificielle comme écriture dynamique. De la raison graphique à la raison computationnelle.” Au nom du sens. Ed. Paolo Fabbri and Jean Petitot. Paris: Grasset, 1999. 290-319. Bacon, Francis. “New Atlantis” [1627]. Lord Bacon’s Works. Vol. 2. Ed. Basil Montagu. London: William Pickering, 1825-34.
Bibliography
519
Bacon, Francis. “New Atlantis.” The Major Works. Ed. Brian Vickers. Oxford: Oxford University Press, 1996. Baecker, Dirk. “Der Nullpunkt.” Brian Rotman. Die Null und das Nichts. Eine Semiotik des Nullpunkts. Berlin: Kadmos, 2000. 7-18. Bailey, F. Lee. “Skull’s Darkroom. The Camera Obscura and Subjectivity.” Philosophy of Technology. Ed. Paul T. Durbin. Dordrecht: Kluwer, 1989. Baldinucci, Filippo. Vocabolario toscano dell’arte del disegno . . . Florence, 1681. Baldus, R. and F. Löbell. Nichteuklidische Geometrie (= Sammlung Göschen, vol. 970/ 970a). Berlin: de Gruyter, 1964. Baltrušaitis, Jurgis. Anamorphoses ou perspectives curieuses. Paris: Perrin, 1955. Baltrušaitis, Jurgis. Anamorphoses ou Magie artificielle des effets merveilleux. Paris: Perrin, 2nd ed.: 1969 Baltrušaitis, Jurgis. Introduction to Freed Leeman, Joost Elffers, and Mike Schuyt. Anamorphosen. Ein Spiel mit der Wahrnehmung, dem Schein und der Wirklichkeit. [Exhibit. cat.] Cologne: DuMont, 1975. Baltrušaitis, Jurgis. Anamorphoses ou Thaumaturgus opticus. Paris: Flammarion, 1996. Bardon, François Dandré. Traité de Peinture suivi d’un essai sur la sculpture pour servir d’introduction à une histoire universelle, relative à ces beaux-arts. 2 vols. Paris, 1765. Baeyer, J.J.. “Bericht über den allgemeinen Standpunkt der Preussischen Vermessungen in Bezug auf die mitteleuropäische Gradmessung, und im Besonderen über die im Jahre 1864 ausgeführten Arbeiten.” General-Bericht über die mitteleuropäische Gradmessung für das Jahr 1864. Berlin: Reimer, 1865. Barkan, Leonard. Unearthing the Past. Archeology and Aesthetics in the Making of Renaissance Culture. New Haven and London: Yale University Press, 1999. Barnett, S.J. Idol Temples and Crafty Priests. The Origins of Enlightenment Anticlericalism. New York: St. Martin’s Press, 1999. Barthes, Roland. Criticism and Truth. Minneapolis: University of Minnesota Press, 1987. Baur-Heinhold, Margarete. Theater des Barock. Festliches Bühnenspiel im 17. und 18. Jahrhundert. München, 1966. Becher, Johann Joachim. Natur=Kuendigung der Metallen. Mit vielen Curiosen/ Beweißthumben/ Natuerlichen Gruenden/ Gleichnuessen/ Erfahrenheiten. Frankfurt, 1661. Becher, Johann Joachim. Experimentum chymicum novum, quo artificialis & instantanea Metallorum Generatio & Transmutatio ad Oculum demonstratur. Frankfurt, 1671. Becher, Johann Joachim. Supplementum secundum in physicam subterraneam, id est . . . Demonstratio Philosophica, seu Theses chymicae, veritatem & possibilitatem transmutationis metallorum in aurum evincentes. Frankfurt, 1675. Becher, Johann Joachim. Trifolium Becherianum Hollandicum oder . . . Dr. Joh. Bechers Drey Neue Erfindungen Bestehende in einer Seiden=Wasser=Muehle und Schmelz= Wercke. Frankfurt, 1679. Becher, Johann Joachim. Chymisches Laboratorium Oder unter-erdische Naturkuendigung. Frankfurt, 1680. Becher, Johann Joachim. Experimentum Chymicum Novum: Oder Neue Chymische Prob, Worinnen Die künstliche gleich=darstellige Transmutation, oder Verwandelung/ derer Metallen/ augenscheinlich dargethan. Frankfurt, 1680. Becher, Johann Joachim. Institutiones chimicae prodromae, id est J.J. Becheri Spirensis Oedipus chimicus. 2nd ed., Frankfurt, 1680. Becher, Johann Joachim. Chymischer Glücks-Hafen oder große Chymische Concordanz und Collection/ Von fuenfzehen hundert Chymischen Processen. Frankfurt, 1682. Becher, Johann Joachim. Tripus hermeticus fatidicus, pandens oracular chymica. Frankfurt, 1689.
520
Bibliography
Becher, Johann Joachim. Chymischer Rosen=Garten, Samt einer Vorrede und kurtz gefassten Lebens=Beschreibung Herrn D. Bechers. Ed. Friedrich Roth-Scholtzen. Nuremberg, 1717. Becker, Howard S. Art Worlds. Berkeley: University of California Press, 1982. Becker, Oskar. Grundlagen der Mathematik in geschichtlicher Entwicklung. Frankfurt a.M.: Suhrkamp, 1975. Beckmann, Johann. A History of Inventions and Discoveries. 4. vols. Trans. William Johnston. London: Longman, Hurst, Rees, Orme, and Brown, 1817. Belaval, Yvon. Leibniz critique de Descartes. Paris: Gallimard, 1960. Belting, Hans. Bild-Anthropologie. Munich: Fink, 2001. Benoist, Jocelyn. L’a priori conceptuel. Paris: Vrin, 1999. Bepler, Jill. “Ansichten eines Staatsbegräbnisses. Funeralwerke und Diarien als Quelle zeremonieller Praxis.” Zeremoniell als höfische Ästhetik in Spätmittelalter und Früher Neuzeit (= Frühe Neuzeit, vol. 25). Ed. Jörg Jochen Berns and Thomas Rahn. Tübingen: Niemeyer, 1995. 183-97. Berger, John. Ways of Seeing. London: BBC/Penguin, 1972. Bergson, Henri. Time and Free Will. An Essay on the Immediate Data of Consciousness. Trans. F.L. Pogson. London: George Allen and Company, 1912. Bergson, Henri. “Laughter.” Comedy. Ed. Wylie Sypher. Baltimore: Johns Hopkins University Press, 1956. Bergson, Henri. The Creative Mind. Trans. Mabelle L. Andison. New York: Greenwood Press, 1968. Bergson, Henri. Creative Evolution. Trans. Arthur Mitchell. New York: Dover, 1998. Bergson, Henri. Matter and Memory. Trans. Nancy Margaret Paul and W. Scott Palmer. New York: Zone Books, 2002. Bergson, Henri, Étienne-Jules Marey et al. “Groupe d’études de phénomènes psychiques (1901).” Mélanges. Ed. André Robinet. Paris: P.U.F, 1972. Berlioz, Hector. Grand traité d’instrumentation et d’orchestration modernes [1843/44]. Paris: Schonenberger, 1855. Bernard, Claude. Introduction to Experimental Medicine [1865]. Trans. Henry Copley Greene. New York: Dover, 1957. Berns, Jörg Jochen. “‘Vergleichung eines Vhrwercks, vnd eines frommen andächtigen Menschens.’ Zum Verhältnis von Mystik und Mechanik bei Spee.” Friedrich von Spee. Dichter, Theologe und Bekämpfer der Hexenprozesse. Ed. Italo Michele Battafarano. Gardolo di Trento: Reverdito, 1988. 101-206. Berns, Jörg Jochen. “Die Jagd auf die Nymphe Echo. Künstliche Echoeffekte in Poesie, Musik und Architektur der Frühen Neuzeit.” Die Mechanik in den Künsten. Studien zur ästhetischen Bedeutung von Naturwissenschaft und Technologie. Ed. idem and Hanno Möbius. Marburg: Jonas Verlag, 1990. 67-82. Berns, Jörg Jochen and Thomas Rahn. “Zeremoniell und Ästhetik.” Zeremoniell als höfische Ästhetik in Spätmittelalter und Früher Neuzeit (= Frühe Neuzeit, vol. 25). Ed. idem. Tübingen: Niemeyer, 1995. 650-66. Berns, Jörg Jochen et al., eds. Erdengötter. Fürst und Hofstaat in der Frühen Neuzeit im Spiegel von Marburger Bibliotheks- und Archivbeständen (= Schriften der UB Marburg, vol. 77). [Exhibit. cat.] Marburg: Universitätsbibliothek, 1997. Berns, Jörg Jochen. “Kriegs- und Friedensbilder. Mittel ihrer ästhetischen Reflexion im 17. Jahrhundert.” Morgen-Glantz 9 (1999): 181-218. Berns, Jörg Jochen and Wolfgang Neuber, eds. Seelenmaschinen. Gattungstraditionen, Funktionen und Leistungsgrenzen der Mnemotechniken vom späten Mittelalter bis zum Beginn der Moderne. Vienna, Cologne, and Weimar: Böhlau, 2000.
Bibliography
521
Bessel, Friedrich Wilhelm. Abhandlungen. 3 vols. Ed. Rudolf Engelmann. Leipzig: Verlag von Wilhelm Engelmann, 1875-1876. Bettinelli, Saverio. Dell’Entusiasmo delle belle arti. Milan, 1769. Betz, Hans Dieter. “Magic and Mystery in the Greek Magical Papyri.” Magika Hiera. Ancient Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 244-98. Beukers, Harm. “Het Laboratorium van Sylvius.” Tijdschrift voor de Geschiedenis van de Geneeskunde, Natuurwetenschappen, Wiskunde en Techniek 3 (1980): 28-36. Bexte, Peter and Werner Künzel. Maschinendenken/Denkmaschinen. An den Schaltstellen zweier Kulturen. Frankfurt a.M.: Insel Verlag, 1996. Biagioli, Mario. Galileo, Courtier. The Practice of Science in the Culture of Absolutism. Chicago and London: University of Chicago Press, 1995. Bigg, Charlotte. “Photography and Labour History of Astrometry. The Carte du Ciel.” The Role of Visual Representations in Astronomy. History and Research Practice. Ed. Klaus Hentschel. Thun and Frankfurt a.M.: Deutsch, 2000. 90-106. Bijleveld, Willem, ed. De Leidse Sterrewacht. Vier eeuwen wacht bij dag en bij nacht. Zwolle: Uitgeverij Waanders/De Kler, 1983. Binnig, Gerd and Heinrich Rohrer. “Scanning Tunneling Microscopy.” Helvetia Physica Acta 55 (1982): 726-35. Binnig, Gerd et al. “Energy-Dependent State-Density Corrugation of a Graphite Surface as Seen by Scanning Tunneling Microscopy.” Europhysics Letters 1 (1986): 31-36. Bjurström, Per. Giacomo Torelli and Baroque Stage Design. Stockholm: Almqvist & Wiksell, 1961. Blondel, Christine and Mathias Dörries, eds. Restaging Coulomb. Usages, controverses et réplications autour de la balance de torsion. Florence: L.S. Olschki, 1994. Blumenberg, Hans. Der Prozeß der theoretischen Neugierde. Frankfurt a.M.: Suhrkamp, 1973. Blumenberg, Hans. “Curiosity is Enrolled in the Catalogue of Vices.” The Legitimacy of the Modern Age. Trans. R.M. Wallace. Cambridge: MIT Press, 1983. 309-25. Blumenberg, Hans. Paradigmen zu einer Metaphorologie. Frankfurt a.M.: Suhrkamp, 1998. Boehm, Gottfried. “Zwischen Auge und Hand. Bilder als Instrumente der Erkenntnis.” Mit dem Auge denken. Strategien der Sichtbarmachung in wissenschaftlichen und virtuellen Welten. Ed. Bettina Heintz and Jörg Huber. Zurich and New York: Springer, 2001. 43-54. Boerhaave, Herman. Boerhaave’s Orations. Ed. and. trans. Elze Kegel-Brinkgreve and Antonie M. Luyendijk-Elshout. Leiden: Brill, 1983. Böger, Ines. ‘Ein Seculum . . . da man zu Societäten Lust hat.’ Darstellung und Analyse der Leibnizschen Sozietätspläne vor dem Hintergrund der europäischen Akademiebewegung im 17. und 18. Jahrhundert. 2 vols. Munich: Utz, 1997. Böhme, Hartmut. “Das Unsichtbare. Mediengeschichtliche Annäherungen an ein Problem neuzeitlicher Wissenschaft.” Performativität und Medialität. Ed. Sybille Krämer. Munich: Fink, 2004. 215-46. Böhme, Hartmut. “The Metaphysics of Phenomena. Telescope and Microscope in the Works of Goethe, Leeuwenhoek, and Hooke.” Collection, Laboratory, Theater. Scenes of Knowledge in 17th Century. Ed. Helmar Schramm, Ludger Schwarte, and Jan Lazardzig. Berlin: de Gruyter, 2005. 355-93. Bolles, Edmund Blair. “Isaac Newton and Robert Hooke, Dispute on the Nature of Light.” Galileo’s Commandment. Ed. idem. New York: W.H. Freeman, 1997. Bonanni, Philippo. Observationes circa viventia, que in rebus non viventibus reperiuntur. Cum Micrographia curiosa. Rome, 1699.
522
Bibliography
Bonnefoy, Yves. Rome, 1630. L’Horizon du premier baroque. Paris: Flammarion, 1994. Bonneuil, Christophe. “The Manufacture of Species. Kew Gardens, the Empire, and the Standardisation of Taxonomic Practices in Late Nineteenth-Century Botany.” Instruments, Travel and Science. Itineraries of Precision from the Seventeenth to the Twentieth Century. Ed. Marie-Noëlle Bourguet, Christian Licoppe, and H. Otto Sibum. London and New York: Routledge, 2002. 189-215. Bordron, Jean-François. “Bergson et les images. L’iconicité de la pensée dans ‘Le possible et le réel.’” Lire Bergson. Le possible et le réel. Ed. Frédéric Cossutta. Paris: P.U.F, 1998. 159-81. Boschini, Marco. La carta del navegar pitoresco. Ed. Anna Pallucchini. Venice: Istituto per la collaborazione culturale, 1966. Bosman, Willem. A New and Accurate Description of the Coast of Guinea. Divided into the Gold, the Slave, and the Ivory Coast. London: Knapton, 1705. Bourguet, Marie-Noëlle. “Landscape with Numbers. Natural History, Travel and Instruments in the Late Eighteenth and Early Nineteenth Centuries.” Instruments, Travel and Science. Itineraries of Precision from the Seventeenth to the Twentieth Century. Ed. idem, Christian Licoppe, and H. Otto Sibum. London and New York: Routledge, 2002. 96-125. Boyle, Robert. The Works of the Honourable Robert Boyle. Ed. Thomas Birch. London, 1772. Boyle, Robert. The Works. Ed. Thomas Birch. Hildesheim: Olms, 1966 [facsimile of the edition London, 1772]. Braarvig, Jens. “Magic. Reconsidering the Grand Dichotomy.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 21-54. Bradley, D.C., G.C. Chang, and R.A. Andersen. “Encoding of Three-Dimensional Structure-from-Motion by Primate Area MT Neurons.” Nature 392 (April 16, 1998): 71417. Brague, Rémi. Wisdom of the World. The Human Experience of the Universe in Western Thought. Trans. Teresa Lavender Fagan. Chicago: University of Chicago Press, 2003. Brain, Robert Michael. The Graphic Method. Inscription, Visualization, and Measurement in Nineteenth-Century Science and Culture. Ph.D. University of California, 1996. Bredekamp, Horst. “Der Mensch als ‘zweiter Gott.’ Motive der Wiederkehr eines kunsttheoretischen Topos im Zeitalter der Bildsimulation.” Interface 1. Elektronische Medien und künstlerische Kreativität. Ed. Klaus Peter Dencker. Hamburg: Verlag HansBredow-Institut für Rundfunk und Fernsehen, 1992. 134-47. Bredekamp, Horst. Antikensehnsucht und Maschinenglauben. Die Geschichte der Kunstkammer und die Zukunft der Kunstgeschichte. Berlin: Wagenbach, 1993. Bredekamp, Horst et al. “Bildwelten des Wissens.” Bilder in Prozessen (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1.1). Ed. idem and Gabriele Werner. Berlin: Akademie Verlag, 2003. 9-20. Bredekamp, Horst. Die Fenster der Monade. Gottfried Wilhelm Leibniz’ Theater der Natur und Kunst. Berlin: Akademie Verlag, 2004. Breidbach, Olaf. Die Materialisierung des Ichs. Zur Geschichte der Hirnforschung im 19. und 20. Jahrhundert. Frankfurt a.M.: Suhrkamp, 1997. Breidbach, Olaf. Das Anschauliche oder über die Anschauung von Welt. Vienna and New York: Springer, 2000. Breidbach, Olaf. Deutungen. Zur philosophischen Dimension der internen Repräsentation. Weilerswist: Velbrück Wissenschaft, 2001.
Bibliography
523
Breidbach, Olaf. “Hirn und Bewußtsein. Überlegungen zu einer Geschichte der Neurowissenschaften.” Neurowissenschaften und Philosophie. Ed. Michael Pauen and Gerhard Roth. Munich: Fink, 2001. 11-58. Breidbach, Olaf. “Oken in der Wissenschaftsgeschichte des 19. Jahrhunderts.” Lorenz Oken (1779-1851). Ein politischer Naturphilosoph. Ed. idem, Hans-Joachim Fliedner, and Klaus Ries. Weimar: Böhlau, 2001. 15-32. Breidbach, Olaf. “The Origin and Development of the Neurosciences.” Theory and Method in the Neurosciences. Ed. Peter K. Machamer, Rick Grush, and Peter McLaughlin. Pittsburgh: University of Pittsburgh Press, 2001. 7-29. Breidbach, Olaf. “Karl Friedrich Burdach.” Naturphilosophie nach Schelling. Ed. Thomas Bach and idem. Stuttgart-Bad Cannstatt: Frommann-Holzboog, 2005. 73-105. Breidbach, Olaf. “On the Representation of Knowledge in Athanasius Kircher.” Collection, Laboratory, Theater. Scenes of Knowledge in the 17th Century. Ed. Helmar Schramm, Ludger Schwarte, and Jan Lazardzig. Berlin and New York: de Gruyter, 2005. 283-302. Breidbach, Olaf and Michael T. Ghiselin. “Lorenz Oken’s ‘Naturphilosophie’ in Jena, Paris, and London.” History and Philosophy of the Life Sciences 24 (2002): 219-47. Brenner, Klaus-Peter. Chipendani und Mbira. Musikinstrumente, nicht-begriffliche Mathematik und die Evolution der harmonischen Progression in der Musik der Shona in Zimbabwe. Göttingen: Vandenhoeck und Ruprecht, 1997. Brian, Éric. “Transactions statistiques au XIXe siècle. Mouvements internationaux de capitaux symboliques.” Actes de la recherche en sciences sociales 145 (décembre 2002): 34-46. Britton, Pierce Dominic. “Lionizing Leonardo. A Physiognomic Conceit in Vasari’s ‘Vite.’” Source 22.4 (2003): 10-15. Bronowski, Jacob. Science and Human Values [1956]. New York: Harper & Row, 1975. Brusatin, Manlio. Storia delle linee. Turin: Einaudi, 1993. Buck, August. “Die Kunst der Verstellung im Zeitalter des Barocks.” Festschrift der Wissenschaftlichen Gesellschaft an der Johann Wolfgang Goethe-Universität Frankfurt a.M. Wiesbaden: Steiner, 1981. 85-103. Budde, Johann Franz. Elementa philosophiae instrumentalis et theoreticae. Vol. 2. Halle, 1703. Bull, Lucien. La cinématographie. Paris: Armand Collin, 1928. Bulmer, Martin. The Chicago School of Sociology. Chicago: University of Chicago Press, 1984. Burckhardt, Martin. Metamorphosen von Raum und Zeit. Eine Geschichte der Wahrnehmung. Frankfurt a.M. and New York: Campus-Verlag, 1994. Burckhardt, Martin. Vom Geist der Machine. Eine Geschichte kultureller Umbrüche. Frankfurt a.M.: Campus, 1999. Burdach, Karl Friedrich. Beyträge zur näheren Kenntnis des Gehirns in Hinsicht auf Physiologie, Medicin und Chirurgie. First Part. Leipzig: Breitkopf und Härtel, 1806. Busche, Hubertus. Leibniz’ Weg ins perspektivische Universum. Eine Harmonie im Zeitalter der Berechnung. Hamburg: Meiner, 1997. Bußmann, Klaus and Heinz Schilling, eds. 1648. Krieg und Frieden in Europa. 3 vols. Munich: Bruckmann, 1998. Butters, Suzanne B. The Triumph of Vulcan. Sculptor’s Tools, Porphyry, and the Prince in Ducal Florence. 2 vols. Florence: Olschki, 1996. Bynum, William F. “The Anatomical Method, Natural Theology, and the Functions of the Brain.” Isis 64 (1973): 445-468. Cahan, David. An Institute for Empire. The Physikalisch-Technische Reichsanstalt, 1871-1918. Cambridge: Cambridge University Press, 1989.
524
Bibliography
Canales, Jimena. “Photogenic Venus. The ‘Cinematographic Turn’ and Its Alternatives in Nineteenth-Century France.” Isis 93 (2002): 585-613. Canales, Jimena. Sensational Differences. Individuality in Observation, Experimentation, and Representation (France 1853-1895). Ph.D. Harvard University, 2003. Canetti, Elias. Crowds and Power. Trans. Carol Stewart. New York: Continnum, 1973. Canguilhem, Georges. La formation du concept de réflexe aux XVIIe et XVIIIe siècles. Paris: P.U.F, 1955. Cannon, Susan Faye. “Humboldtian Science.” Science in Culture. The Early Victorian Period. New York: Science History Publications, 1978. Capivaccius, Hieronymus. Opera Omnia. Ed. Johannes-Hartmannus Beyerus. Frankfurt, 1603. Carabelli, Giancarlo. In the Image of Priapus. London: Duckworth, 1996. Carr, William. An Accurate Description of the United Netherlands, And of the Most Considerable Parts of Germany, Sweden, and Denmark. London, 1691. Carruthers, Mary. The Book of Memory. A Study of Memory in Medieval Culture. Cambridge: Cambridge University Press, 1992. Carruthers, Mary. The Craft of Thought. Meditation, Rhetoric, and Making of Images. Cambridge: Cambridge University Press, 2000. Castiglione, Baldassare. Il libro del cortegiano. Ed. Walter Barberis. Turin: Einaudi, 1998. Caus, Salomon de. Von gewaltsamen Bewegungen . . . Hannover: Vincentz, 1977 [facsimile of the edition Frankfurt, 1615]. Cavanaugh, Maureen B. Eleusis and Athens. Documents in Finance, Religion, and Politics in the Fifth Century B.C. Atlanta: Scholars Press, 1996. Cavendish, Henry. “An Account of Some Attempts to Imitate the Effects of the Torpedo by Electricity.” Philosophical Transactions of the Royal Society 66 (1776): 196-225. Centralbureau der Europäischen Gradmessung. Verhandlungen der vom 13. bis 16. September 1880 zu München abgehaltenen sechsten allgemeinen Conferenz der Europäischen Gradmessung. Berlin: Reimer, 1881. Cernuschi, Claude. ‘Not an Illustration, but the Equivalent.’ A Cognitive Approach to Abstract Expressionism. Cranbury: Associated University Presses, 1997. Chadarevian, Soraya de. Designs for Life. Cambridge: Cambridge University Press, 2002. Chardère, Bernard. Les images des Lumière. Paris: Gallimard, 1995. Chastel, André, ed. Leonardo da Vinci. Sämtliche Gemälde und die Schriften zur Malerei. Munich: Schirmer/Mosel, 1990. Chevalley, Catherine. “L’ars magna lucis et umbrae d’Athanase Kircher. Néoplatonisme, hermétisme et ‘nouvelle philosophie.’” Baroque 12 (1989): 95-109. Choulant, Ludwig. History and Bibliography of Anatomic Illustration. Trans. Mortimer Frank. New York: Hafner, 1945 Christie, James. A Disquistion upon Etruscan Vases. Displaying their Probable Connection with the Shows at Eleusis, and the Chinese Feast of Lanterns, with Explanations of a few of the Principal Allegories Depicted upon them. London: T. Becket. 1806. Christie, James. An Essay on That Earliest Species of Idolatry. The Worship of the Elements. Norwich: Stanhopian Press, 1814. Cipolla, Carlo M. Money, Prices, and Civilization in the Mediterranean World. Princeton: Princeton University Press, 1956. Clark, Andy. Being There. Putting Brain, Body and World Together Again. Cambridge: MIT Press, 1997. Clark, Kenneth. The Drawings of Leonardo da Vinci in the Collection of H. M. The Queen at Windsor Castle. 3 vols. London: Phaidon, 1968.
Bibliography
525
Clarke, Edwin and Kenneth Dewhurst. An Ilustrated History of Brain Function. Berkeley: University of California Press, 1972. Clercq, Peter de. “Exporting Scientific Instruments around 1700. The Musschenbroek Documents at Marburg.” Tractrix 3 (1991): 79-120. Clercq, Peter de. At the Sign of the Oriental Lamp. The Musschenbroek Workshop in Leiden, 1660-1750. Rotterdam: Erasmus Publications, 1997. Clerk, John. Memoirs of the Life of Sir John Clerk of Penicuik. Extracted by Himself from His Own Journals 1676-1755. Ed. John M. Gray. Edinburgh: T.&A., 1892. Coiter, Volcher. Externarum et internarum principalium humani corporis partium tabulae atque anatomicae exercitationes observationesque variae. Nuremberg, 1572. Cole, Francis Joseph. A History of Comparative Anatomy. From Aristotle to the Eighteenth Century. London: Macmillan, 1944. Collins, Harry. “Tacit Knowledge and Scientific Networks.” Science in Context. Readings in the Sociology of Science. Ed. Barry Barnes and David Edge. Milton Keynes: Open University Press, 1982. 44-64. Colombo, Realdo. De re anatomice libri XV/XVI. Venice, 1559. Conwell, Joseph F. Contemplation in Action. A Study in Ignatian Prayer. Spokane: Gonzaga University Press. 1957. Cook, Harold J. “Physick and Natural History in Seventeenth-Century England.” Revolution and Continuity. Essays in the History and Philosophy of Early Modern Science. Ed. Peter Barker and Roger Ariew. Washington: Catholic University of America Press, 1991. 63-80. Cooper, Alix. “The Household.” The Cambridge History of Early Modern Science. Ed. Katharine Park and Lorraine Daston. Cambridge: Cambridge University Press, 2006. 224-237. Corbin, Alain. Village Bells. The Culture of the Senses in Nineteenth-Century French Countryside. Trans. Martin Thom. New York: Columbia University Press, 1998. Corneille, Pierre. Andromède. Tragédie. Ed. Cristian Delmas. Paris: Didier, 1974. Corneille, Thomas. Circè, tragédie, ornée de machines, de changemens de théâtre, & de musique. Paris, 1690. Costamagna, Philippe. Pontormo. Milan: Electa, 1994. Cotreau, James D. The Historical Development of the Universal Postal Union and the Question of Membership. Boston: n. publ., 1975. Creager, Angela N.H. The Life of a Virus. Tobacco Mosaic Virus as an Experimental Model. Chicago: University of Chicago Press, 2002. Cressey, Paul G. The Taxi-Dance Hall. Chicago: University of Chicago Press, 1932. Cronon, William. Nature’s Metropolis. Chicago and the Great West. New York: Norton, 1991. Crosby, Alfred W. The Measure of Reality. Quantification and Western Society 12501600. Cambridge: Cambridge University Press, 1997. Curl, James Stevens. The Art and Architecture of Freemasonry. An Introductory Study. London: B.T. Batsford, 1991. D’Eramo, Marco. The Pig and the Skyscraper. Chicago, A History of Our Future. London: Verso, 2002. Damisch, Hubert. L’origine de la perspective. Paris: Flammarion, 1987. Daston, Lorraine. Eine kurze Geschichte der wissenschaftlichen Aufmerksamkeit. Munich: C.F. von Siemens Stiftung, 2000. Daston, Lorraine, ed. The Biography of Scientific Objects. Chicago: University of Chicago Press, 2000. Daston, Lorraine and Peter Galison. “The Image of Objectivity.” Representations 40 (1992): 81-128.
526
Bibliography
Daston, Lorraine and Peter Galison. “Das Bild der Objektivität.” Ordnungen der Sichtbarkeit. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 29-99. Daston, Lorraine and Katharine Park. Wonders and the Order of Nature, 1150-1750. New York: Zone Books, 1998. David, F.A. Antiquites etrusques, grecques et romaines. Gravees par . . . avec les explications par [Baron] d’Hancarville. Paris: Chez l’Auteur, 1785-1787. Davis, Mike. Ecology of Fear. Los Angeles and the Imagination of Disaster. New York: Random House, 1998. De Sepi, Giorgio. Musaeum celeberrimum. Amsterdam, 1678. Débarbat, Suzanne et al., eds. Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Dordrecht, Boston, and London: Kluwer, 1988. Deegan, Mary Jo. Jane Addams and the Men of the Chicago School, 1892-1918. New Brunswick: Transaction Books, 1988. Degenhart, Bernhard. “Zur Graphologie der Handzeichnung. Die Strichbildung als stetige Erscheinung innerhalb der italienischen Kunstkreise.” Kunstgeschichtliches Jahrbuch der Bibliotheca Hertziana 1 (1937): 223-343. Delavigne, V. and M. “Un grand Seigneur au XVIIe siècle. Le Marquis de Sourdéac.” La Revue Hebdomadaire 20 (1911): 450-83. Deleuze, Gilles. The Movement-Image. Cinema 1. Trans. Hugh Tomlinson and Barbara Habberjam. London: Athlone Press, 1986. Deleuze, Gilles. The Time-Image. Cinema 2. Trans. Hugh Tomlinson and Robert Galeta. London: Athlone Press, 1989. Deleuze, Gilles. The Fold. Leibniz and the Baroque. Trans. Tom Conley. Minneapolis: University of Minnesota Press, 1992. Deleuze, Gilles and Félix Guattari. What is Philosophy? New York: Columbia University Press, 1994. Denicke, C.L. Vollständiges Lehrgebäude der ganzen Optik oder der Sehe-, Spiegel- und Strahlbrech-Kunst . . . Altona, 1757. Der Triumphzug Kaiser Maximilians I. 1516-1518. 147 Holzschnitte von Albrecht Altdorfer, Hans Burgkmair, Albrecht Dürer u.a. Mit dem von Kaiser Maximilian diktierten Programm (= Harenberg Kommunikation, vol. 100). Afterword by Horst Appuhn. Dortmund: Harenberg Kommunikation, 1979. Descartes, René. De Homine. Leiden, 1662. Descartes, René. Traité de la Mechanique . . . Paris, 1668. Descartes, René. “Discours de la méthode” [1637]. Œuvres de Descartes. Vol. 6. Ed. Charles Adam and Paul Tannery. Paris: Léopold Cerf, 1902. 1-78. Descartes, René. Œuvres de Descartes. 13 vols. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1969 Descartes, René. Geometrie. Ed. Ludwig Schlesinger. Darmstadt: Wissenschaftliche Buchgesellschaft, 1981. Descartes, René. “Discourse on the Method.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdach. Cambridge: Cambridge University Press, 1984. Vol. 1, 111-175. Descartes, René. “Meditations on First Philosophy.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdach. Cambridge: Cambridge University Press, 1984. Vol. 2, 1-62. Descartes, René. “The Passions of the Soul.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. 1, 325-404.
Bibliography
527
Descartes, René. “The Treatise on Man.” The Philosophical Writings of Descartes. 3 vols. Trans. John Cottingham, Robert Stoothoff, and Dugald Murdoch. Cambridge: Cambridge University Press, 1985. Vol. 1, 99-108. Descartes, René. Œuvres de Descartes. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1996. Descartes, René. The World and Other Writings. Ed. Stephen Gaukroger. Cambridge: Cambridge University Press, 1998. Descartes, René. Œuvres complètes. Ed. Charles Adam and Paul Tannery. Paris: Vrin, 1999. Dettelbach, Michael. “Global Physics and Aesthetic Empire. Humboldt’s Physical Portrait of the Tropics.” Visions of Empire. Voyages, Botany, and Representations of Nature. Ed. Peter H. Reill and David Philip Miller. Cambridge: Cambridge University Press, 1996. 258-92. Dewhurst, Kenneth. Willis’s Oxford Lectures. Oxford: Sandford Publications, 1980. Dewitz, Bodo von and Werner Nekes. Ich sehe was, was Du nicht siehst! Sehmaschinen und Bilderwelten. Die Sammlung Werner Nekes. Göttingen: Steidl, 2002. Dickie, Matthew W. “The Learned Magician and the Collection and Transmission of Magical Lore.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 163-93. Diderot, Denis. “Lettre à Landois du 29 juin 1756.” Correspondance. Paris: Éditions de Minuit, 1955. Vol. 1, 209. Diderot, Denis. Thoughts on the Interpretation of Nature and other Philosophical Works. Intro. David Adams. Trans. Lorna Sandler. Manchester: Clinamen Press, 1999. Diderot, Denis and Jean LeRond d’Alembert. Encyclopédie, ou Dictionnaire raisonné des sciences, des arts et des métiers, par une société de gens de lettres. 35 vols. Ed. idem. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 1751-1780. Diderot, Denis. “Eclectisme.” Encyclopédie, ou Dictionnaire raisonné des sciences, des arts et des métiers par une société de gens de lettres. 35 vols. Ed. idem and Jean LeRond d’Alembert. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 17511780. Vol 5, 270-93. Diderot, Denis and Jean LeRond d’Alembert. “Eclipse.” Encyclopédie, ou Dictionnaire raisonné des sciences, des arts et des métiers, par une société de gens de lettres. 35 vols. Ed. idem. Paris: Chez Briasson, David l’Aine, Le Breton, Durand, 17511780. Vol 5, 293-98. Didi-Huberman, Georges. Ce que nous voyons, ce qui nous regarde. Paris: Éditions de Minuit, 1992. Didi-Huberman, Georges. “L’image-sillage.” L’Inactuel 10 (2004): 111-126. Didi-Huberman, Georges. Confronting Images. Questioning the Ends of a Certain History of Art. Pennsylvania: Pennsylvania State University Press, 2005. Diedrichs, Christof. Vom Glauben zum Sehen. Die Sichtbarkeit der Reliquie im Reliquiar. Ein Beitrag zur Geschichte des Sehens. Berlin: Weißensee-Verlag, 2001. Dierig, Sven. Wissenschaft in der Maschinenstadt. Emil Du Bois-Reymond und seine Laboratorien in Berlin. Göttingen: Wallenstein, 2006. Dietrich, Margret. “Der Barocke Corneille. Ein Beitrag zum Maschinen-Theater des 17. Jahrhunderts.” Maske und Kothurn 4 (1958): 199-219, 316-45. Dietrich, Margret. “Vom Einfluss der Mathematik and Mechanik auf das Barocktheater.” Sonderabdruck aus dem Anzeiger der phil.-hist. Klasse der Österreichischen Akademie der Wissenschaften 107.1 (1970): 7-22. Diner, Steven J. “Department and Discipline. The Department of Sociology at the University of Chicago, 1892-1920.” Minerva 13 (1975): 514-53.
528
Bibliography
Domenichi, Ludovico. L’idea del theatro dell’ecc. M. Giulio Camillo. Florenz, 1550. Dominici, Bernardo di. Vite de’ Pittori, Scultori ed Architetti napoletani. Naples, 174245. Dörries, Matthias. “Balances, Spectroscopes, and the Reflexive Nature of Experiments.” Studies in the History and Philosophy of Science 25 (1994): 1-36. Dryden, John. The Works. 20 vols. Ed. Edward Niles Hooker. Berkeley, Los Angeles, and London: University of California Press, 1971. Du Halde, J.B. Description géographique, historique, chronique, politique, et physique de l’empire de la Chine et de la Tartarie chinoise. Paris: P.G. Lemercier, ImprimeurLibraire, 1735. DuBos, Jean Baptiste. Réflexions critiques sur la poésie et la peinture. Paris, 1719. Eamon, William. Science and the Secrets of Nature. Books of Secrets in Medieval and Early Modern Culture. Princeton: Princeton University Press, 1994. Eckermann, Johann Peter. Gespräche mit Goethe in den letzten Jahren seines Lebens. Berlin: Aufbau-Verlag, 1982. Eitrem, Samson. “Dreams and Divination in Magical Ritual.” Magika Hiera. Ancient Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 175-88. Elias, Norbert. The Court Society. Oxford: Basil Blackwell, 1983. Elzinga, Aant and Catharina Landström, eds. Internationalism in Science. London: Taylor Graham, 1996. Encyclopedia Britannica. Vol. 4. New York: Encyclopedia Britannica, 1929. Ennenbach, Wilhelm. “Über eine öffentliche Einrichtung zur Vorführung, Lagerung und Erfassung technischer Objekte.” Neue Museumskunde 24.2 (1981): 103-08. Erxleben, Johann Christian Polykarp. Anfangsgründe der Naturlehre. 6th ed. Göttingen, 1794. Eser, Thomas. Schiefe Bilder. Die Zimmernsche Anamorphose und andere Augenspiele aus den Sammlungen des Germanischen Nationalmuseums. Nuremberg: Germanisches Nationalmuseum, 1998. Evelyn, John. The Diary. 6 vols. Oxford: Clarendon Press, 1955. Fabri, Mario and Elvira Zorzi, eds. Il luogo teatrale a Firenze. Milano: Electa ed., 1975. Fabricius ab Aquapendente, Hieronymus. Opera omnia anatomica. Leipzig, 1687. Fahrenheit, Daniel Gabriel. Fahrenheit’s Letters to Leibniz and Boerhaave. Ed. and trans. Pieter van der Star. Amsterdam: Rodopi, 1983. Farago, Claire J. Leonardo da Vinci’s ‘Paragone.’ A Critical Interpretation with a New Edition of the Text of the ‘Codex Urbinas.’ Leiden: Brill, 1992. Faraone, Christopher A. “The Agonistic Context of Early Greek Binding Spells.” Magika Hiera. Ancient Greek Magic and Religion. Ed. Christopher A. Faraone and Dirk Obbink. New York and Oxford: Oxford University Press, 1991. 3-32. Faris, Robert E.L. Chicago Sociology 1920-1932. Chicago: University of Chicago Press, 1967. Fine, Gary Alan, ed. A Second Chicago School? Chicago: University of Chicago Press, 1995. Faye, Hervé. “Rapport sur le rôle de la photographie dans l’observation du passage de Vénus.” Recueil des mémoires, rapports et documents relatifs à l’observation du passage de Vénus sur le soleil. Paris: Firmin Didot Frères, 1874. 227-36. Faye, Hervé. “Sur l’observation photographique des passages de Vénus et sur un appareil de M. Laussedat.” Recueil des mémoires, rapports et documents relatifs à l’observation du passage de Vénus sur le soleil. Paris: Firmin Didot Frères, 1874. 175-85. Fehrenbach, Frank. Licht und Wasser. Zur Dynamik naturphilosophischer Leitbilder im Werk Leonardo da Vincis. Tübingen: Wasmuth, 1997.
Bibliography
529
Fehrenbach, Frank. “Hier stehe ich, aber mein Auge durcheilt die Räume. Die Vogelstudien und Flugversuche von Leonardo da Vinci.” Frankfurter Allgemeine Zeitung (April 27, 2002): 47. Fehrenbach, Frank. “Komposition.” Metzler Lexikon Kunstwissenschaft. Ideen, Methoden, Begriffe. Ed. Ulrich Pfisterer. Stuttgart and Weimar: Metzler, 2003. 178-83. Félibien, André. Rélation de la fête de Versailles du 1668 – Les Divertissements de Versailles. Paris: Dédale, Maisonneuve et Larose, 1994. Fick-Michel, Nicole. Art et mystique dans les Métamorphoses d’Apulée. Paris: Diffusion Les Belles-Lettres, 1991. Findlen, Paula. Possessing Nature. Museums, Collecting, and Scientific Culture in Early Modern Italy. Berkeley and Los Angeles: University of California Press, 1994. Findlen, Paula. “Masculine Prerogatives. Gender, Space, and Knowledge in the Early Modern Museum.” The Architecture of Science. Ed. Peter Galison and Emily Thompson. Cambridge and London: MIT Press, 1999. 29-58. Fink, Hans-Werner. “Mono-atomic Tips for Scanning Tunneling Microscopy.” IBM Journal of Research and Development 30.5 (1986): 460-65. Fischer, Heinz-Joachim. Der heilige Kampf. Geschichte und Gegenwart der Jesuiten. Munich: Piper, 1987. Flammarion, Camille. “La photographie céleste à l’Observatoire de Paris.” Revue d'Astronomie Populaire 5 (1886): 42-57. Flammarion, Camille. “Le Congrès astronomique pour la photographie du ciel.” Astronomie 6 (1887): 161-69. Fleckner, Uwe. “‘Pourquoi une belle esquisse nous plaît-elle plus qu’un tableau?’ Fragonard, Diderot et l’éloquence du pinceau dans quelques portraits du XVIIIe siècle.” L'art et les normes sociales au XVIIIe siècle. Ed. Thomas W. Gaehtgens. Paris: Editions de la Maison des sciences de l'homme, 2001. 509-33. Fleming, James Rodger. Meteorology in America, 1800-1870. Baltimore and London: Johns Hopkins University Press, 1990. Fletcher, Richard. The Conversion. From Paganism to Christianity. New York: Henry Holt and Company, 1997. Fludd, Robert. Utriusque cosmi maioris scilicet et minoris metaphysica, physica atque technica historia in duo volumina secundum cosmi differentiam divisa. Tomus primus. De macrocosmi historia in duos tractatus divisa. Oppenheim, 1617. Fludd, Robert. Microcosmi historia. Tomus secundus de supernaturali, naturali et contranaturali microcosmi historia in tractatus tres distributa. Oppenheim, 1619. Fludd, Robert. Anatomiae amphitheatrum effigie triplici, more et conditione varia designatum. Frankfurt, 1623. Fontenelle, Bernard Le Boyer de. “Traité de la Liberté.” Textes choisis. Ed. Maurice Roelens. Paris: Editions sociales, 1967. Fontenelle, Bernard Le Bovier de. Conversations on the Plurality of Worlds. Trans. H.A. Hargreaves. Berkeley and Los Angeles: University of California Press, 1990. Forbes, George. The Transit of Venus. London and New York: Macmillan, 1874. Franke, Birgit. “Automaten in höfischen Lustgärten in der Frühen Neuzeit.” Automaten in Kunst und Literatur des Mittelalters und der Frühen Neuzeit (= Wolfenbütteler Mittelalter-Studien, vol. 17). Ed. Klaus Grubmüller and Markus Stock. Wiesbaden: Harrassowitz, 2003. 247-67. French, Roger. William Harvey´s Natural Philosophy. Cambridge: Cambridge University Press, 1994. Freundlich, Francis. Le monde du jeu à Paris 1715-1800. Paris: Albin Michel, 1995. Friedman, Robert Marc. Appropriating the Weather. Vilhelm Bjerknes and the Construction of a Modern Meteorology. Ithaca: Cornell University Press, 1989.
530
Bibliography
Frizot, Michel. Étienne-Jules Marey. Chronophotographie. Paris: Nathan/Delpire, 2001. Fuchs, Thomas. Die Mechanisierung des Herzens. Frankfurt a.M.: Suhrkamp, 1992. Furetière, Antoine. Dictionnaire universel . . . 3 vols. Den Haag and Rotterdam, 1690. Furttenbach, Joseph. Mechanische Reißladen . . . Augsburg, 1644. Galilei, Galileo. Dialogue Concerning the Two Chief World Systems. Trans. Stillman Drake. Berkeley: University of California Press, 1953. Galison, Peter. Image and Logic. A Material Culture of Microphysics. Chicago and London: University of Chicago Press, 1997. Galison, Peter. Einstein’s Clocks, Poincaré’s Maps. Empires of Time. New York: W.W. Norton, 2003. Galison, Peter and Emily Thompson, eds. The Architecture of Science. Cambridge: MIT Press, 1999. Galluzzi, Paolo, ed. Prima di Leonardo. Cultura delle macchine a Siena nel Rinascimento. Milan: Electa, 1991. Galluzzi, Paolo. Gli ingegneri del Rinascimento da Brunelleschi a Leonardo da Vinci. Florence: Giunti, 2001. Gandelman, Claude. “Der Gestus des Zeigens.” Der Betrachter ist im Bild. Kunstwissenschaft und Rezeptionsästhetik. Ed. Wolfgang Kemp. Berlin: Reimer, 1992. 71-93. Gehler, D. Johann Samuel Traugott. Physikalisches Wörterbuch. 11 vols. Leipzig, 18251845. Geimer, Peter, ed. Untot. Verhältnisse von Leben und Leblosigkeit. Preprint No. 250. Berlin: Max-Planck-Institut für Wissenschaftsgeschichte, 2002. Geimer, Peter. “Was ist kein Bild? Zur ‘Störung der Verweisung.’” Ordnungen der Sichtbarkeit. Fotografie in Wissenschaft, Technologie und Kunst. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 313-41. Geimer, Peter. “Weniger Schönheit. Mehr Unordnung. Eine Zwischenbemerkung zu ‘Wissenschaft und Kunst.’” Neue Rundschau 114.3 (2003): 26-38. General-Bericht über die mitteleuropäische Gradmessung für das Jahr 1862. Berlin: Reimer, 1863. Genz, Henning. Gedankenexperimente. Weinheim: Wiley-VCH, 1999. Georges, Heinrich. Lateinisch-Deutsches Handwörterbuch. Hanover, 1892. Ghiselin, Michael T. “Lorenz Oken.” Naturphilosophie nach Schelling. Ed. Thomas Bach and Olaf Breidbach. Stuttgart-Bad Cannstatt: Frommann-Holzboog, 2005. Gibson, Ann Eden. Abstract Expressionism. Other Politics. New Haven: Yale University Press, 1997. Gieryn, Thomas F. “Three Truth-Spots.” Journal of the History of the Behavioral Sciences 38 (2002): 113-32. Gilpin, William. Observations on the Highlands of Scotland. Richmond: Richmond Publishing Co., 1973 [facsimile of the edition London, 1789]. Gipper, Andreas. Wunderbare Wissenschaft. Literarische Strategien naturwissenschaftlicher Vulgarisierung in Frankreich. Munich: Fink, 2002. Goethe, Johann Wolfgang von. Italian Journey (1786-1788). Trans. W.H. Auden and Elizabeth Mayer. San Francisco: North Point Press, 1982. Goldstein, E.B. Sensation and Perception. Pacific Grove: Brooks/Cole, 1999. Gomez de Liano, Ignacio. Athanasius Kircher (1602-1680), Itinerario del extasis, o las imagines de un saber universal. Madrid: Ediciones Siruela, 2001. Gonnessiat, François. Recherches sur l’équation personnelle dans les observations astronomiques de passage. Paris: G. Masson, 1892. Goodman, Nelson. Languages of Art. Indianapolis: Hacket, 1976. Goody, Jack. The Domestication of the Savage Mind. Cambridge: Cambridge University Press, 1977.
Bibliography
531
Goody, Jack. The Logic of Writing and the Organization of Society. Cambridge: Cambridge University Press, 1986. Graf, Fritz. La magie dans l’antiquité greco-romaine. Idéologie et pratique. Paris: Les Belles Lettres, 1994. Graf, Fritz. “Magic and Divinations.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. 283-98. Grafton, Anthony. “Civic Humanism and Scientific Scholarship at Leiden.” The University and the City. From Medieval Origins to the Present. Ed. Thomas Bender. Oxford: Oxford University Press, 1988. 59-78. Griendel, Johann Frantz. Micrographia nova. Nuremberg, 1687. Grmek, Mirko. La première révolution biologique, Réflexions sur la physiologie et la médecine du XVIIe siècle. Paris: Payot, 1990. Guarino, Raimondo. La Tragedia e le macchine. Andromède di Corneille e Torelli. Rome: Bulzoni, 1982. Guckenberger, Reinhard. “Test Structures and Calibration.” Procedures in Scanning Probe Microscopies. Ed. Richard J. Colton et al. Chichester: Wiley, 1998. 9-21. Guilbaut, Serge. How New York Stole the Idea of Modern Art. Chicago: University of Chicago Press, 1983. Haack, Wolfgang. Darstellende Geometrie (= Sammlung Göschen, vol. 144). Berlin: de Gruyter, 1969. Haase, Rudolf. Johann Keplers Weltharmonik. Der Mensch im Geflecht von Musik, Mathematik und Astronomie. Munich: Diederichs, 1998. Habrich, Christa. Loch im Kopf. Zur Geschichte der Schädeltrepanation (= Sammelblätter des Deutschen Medizinhistorischen Museums). Ingolstadt, 2002. Hagner, Michael. “Das Ende vom Seelenorgan: Über einige Beziehungen von Philosophie und Anatomie im frühen 19. Jahrhundert.” Das Gehirn – Organ der Seele? Zur Ideengeschichte der Neurobiologie. Ed. Ernst Florey and Olaf Breidbach. Berlin: Berlin Verlag, 1993. Hahn, Roger. The Anatomy of a Scientific Institution. The Paris Academy of Sciences, 1666-1803. Berkeley and Los Angeles: University of California Press, 1971. Haines, Margaret. “Brunelleschi and Bureaucracy: The Tradition of Public Patronage at the Florentine Cathedral.” I Tatti Studies 3 (1989): 89-125. Haller, Albrecht von. Elementa Physiologiae corporis humani. Vol. 4. Lausannes, 1757-66. Hallyn, Fernand. “Aspects de la problématique de l’illusion chez Descartes.” L’illusion au XVIIe siècle (= Littératures classiques, vol. 44). Ed. Patrick Dandrey and Georges Forestier. Paris: Champion, 2002. 284-304. Hambly, Maja. Drawing Instruments. 1580-1980. London: Wilson, 1988. Hamill, John. The Craft. A History of English Freemasonry. Wellingsborough, Northamptonshire: Crucible, 1986. Hammerstein, Reinhold. Die Musik der Engel. Untersuchungen zur Musikanschauung des Mittelalters. Bern and Munich: Francke, 1962. Hammerstein, Reinhold. Macht und Klang. Tönende Automaten als Realität und Fiktion in der alten und mittelalterlichen Welt. Bern: Francke, 1986. Hannaway, Owen. “Laboratory Design and the Aim of Science. Andreas Libavius versus Tycho Brahe.” Isis 77 (1986): 585-610. Hanslick, Eduard. On the Musically. Beautiful. A Contribution towards the Revision of the Aesthetics of Music. Ed. and trans. Geoffrey Payzant. Indianapolis: Hackett, 1986. Hartung, Gerald. “Johann Joachim Becher oder: Die Projekte und Konzepte eines närrischen Gelehrten.” Zwischen Narretei und Weisheit. Biographische Skizzen und Kon-
532
Bibliography
turen alter Gelehrsamkeit. Ed. Gerald Hartung and Wolf Peter Klein. Hildesheim: Olms, 1997. 262-87. Hartung, Gerald. “Johann Joachim Becher.” Die Philosophie des 17. Jahrhunderts. (= Grundriss der Geschichte der Philosophie). Vol. 4.1. Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Basel: Schwabe, 2001. 190-99. Harvey, William. Exercitatio Anatomica de Motu Cordis et Sanguinis in Animalibus. Frankfurt a.M., 1628. Harvey. William. The Works of William Harvey. Ed. and trans. Robert Willis. London: Sydenham Society, 1847. Harvey, William. De Motu Locali Animalium. Ed. and trans. Gweneth Whitteridge. Cambridge: Cambridge University Press, 1959. Haugeland, John, ed. Mind Design. Cambridge: MIT Press, 1981. Hauser, Arnold. Sozialgeschichte der Kunst und Literatur. Munich: C.H. Beck, 1969. Hayek, Friedrich von. Die Entnationalisierung des Geldes. Tübingen: Mohn, 1977. Heim, Michael. Electric Language. A Philosophical Study of Word Processing. New Haven: Yale University Press, 1987. Heintz, Bettina and Jörg Huber. “Der verführerische Blick.” Mit dem Auge denken. Ed. idem. Zurich and Vienna: Springer, 2001. Helden, Anne C. van. “The Age of the Air-Pump.” Tractrix 3 (1991): 149-72. Hellmann, Gustav. “Die Entwicklung der meteorologischen Beobachtungen in Deutschland von den ersten Anfängen bis zur Einrichtung staatlicher Beobachtungsnetze.” Abhandlungen der Preussischen Akademie der Wissenschaften, Physikalisch-Mathematische Klasse 1 (1926): 1-25. Hennig, Jochen. “Vom Experiment zur Utopie. Bilder in der Nanotechnologie.” Instrumente des Sehens (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, Bd. 2,2). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag. 9-18. Henrich, Dieter. “Hegels Logik der Reflexion.” Die Wissenschaft der Logik und die Logik der Reflexion. Ed. idem. Bonn: Bouvier, 1978. 203-324. Hertel, Christian Gottlieb. Vollständige Anweisung zum Glass-Schleiffen . . . Halle, 1716. Hertz, John, Anders Krogh, and Richard G. Palmer. Introduction to the Theory of Neural Computation. City Redwood: Perseus Books Group, 1991. Hilbert, David. “Über das Unendliche.” Mathematische Annalen 95 (1926): 161-90. Hirsch, Adolphe. “Sur les corrections et équations personnelles dans les observations chronographiques de passage.” Bulletin de la Société des sciences naturelles de Neuchâtel 6 (1863): 365-72. Hirschberg, Walter. “Early Historical Illustrations of West and Central African Music.” African Music 4.3 (1969): 6-18. Hobbes, Thomas. “Elements of Philosophy. The first section, concerning the body (De corpore).” The English Works. Ed. William Molesworth. London: Bohn, 1839. Hobbes, Thomas. Leviathan. London: Dent, 1934. Hobbes, Thomas. Leviathan. Middlesex: Penguin Books, 1968. Hockney, David. Secret Knowledge. Rediscovering the Lost Techniques of the Old Masters. New York: Putnam, 2001. Hoffleit, Dorrit. Some Firsts in Astronomical Photography. Cambridge: Harvard College Observatory, 1950. Hoffmann, Friedrich. Fundamenta Medicinae ex principiis naturae mechanicis in usu Philiatrorum succinte proposita. Halle, 1695. Holden, Constance. “The Origin of Speech.” Science 303 (February 27, 2004): 1316-19. Holländer, Hans. “Anamorphotische Perspektiven und Cartesianische Ornamente. Zu einigen Gemälden von Jean-François Nicéron.” Rezeption und Produktion zwischen
Bibliography
533
1570 und 1730. Festschrift für Günther Weydt. Ed. Wolfdietrich Rasch, Hans Geulen, and Klaus Haberkamm. Munich and Bern: Francke, 1972. 53-76. Holländer, Hans. “Mathematisch-mechanische Capriccios.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 347-54. Holländer, Hans. “Spielformen der Mathesis universalis.” Erkenntnis – Erfindung – Konstruktion. Studien zur Bildgeschichte von Naturwissenschaften und Technik vom 16. bis zum 19. Jahrhundert. Ed. idem. Berlin: Mann, 2000. 325-345. Holsboer, S. Wilma. L’histoire de la mise en scène dans le théâtre français de 1600 à 1657. Paris: Droz, 1933. Hoog, Adriaan C. de. Some Currents of Thought in Dutch Natural Philosophy, 16751720. Ph.D. Oxford, 1974. Hooke, Robert. Micrographia or Some Physiological Descriptions of Minute Bodies Made by Magnifying Glasses with Observations and Inquiries there upon. London, 1665. Hooke, Robert. Micrographia or some Physiological Descriptions of Minute Bodies made by Magnifying Glasses. New York: Dover, 1961 [facsimile of the edition London, 1665]. Hooke, Robert. The Posthumous Works. New York and London: Johnson, 1969 [facsimile of the edition London, 1705]. Horkheimer, Max. Critique of Instrumental Reason. Trans. Matthew J. O’Connell et al. New York: Continuum, 1974. Horkheimer, Max. Eclipse of Reason. New York: Continuum, 1974. Horkheimer, Max and Theodor W. Adorno. Dialectic of Enlightenment. Philosophical Fragments. Ed. Gunzelin Schmid Noerr. Trans. Edmund Jephcott. Stanford: Stanford University Press, 2002. Horn, Angelica. “Das Experiment der Zentralperspektive. Filippo Brunelleschi und René Descartes.” Descartes im Diskurs der Neuzeit. Ed. Wilhelm Friedrich Niebel and Herbert Schnädelbach. Frankfurt a.M.: Suhrkamp, 2000. 9-32. Horn-Monval, Madeleine. “La grande machinerie théâtrale et ses origines.” Revue d’histoire du théâtre 4 (1958): 291-308. Horstmann, Rolf-Peter. Ontologie und Relationen. Bradley, Russell und die Kontroverse über interne und externe Beziehungen. Königstein: Athenaeum, 1984. Hoyt, Homer. One Hundred Years of Land Values in Chicago. Chicago: University of Chicago Press, 1933. Huarte de San Juan, Juan. Examen de ingenios para las ciencias. Baeza, 1575. Hughes, J. Trevor. Thomas Willis. London: Royal Society of Medicine Services Limited, 1991. Hunter, Michael. The Royal Society and Its Fellows 1660-1700. Morphology of an Early Scientific Institution. Chalfont St. Giles: British Society for the History of Science, 1982. Husserl, Edmund. Logische Untersuchungen. Ed. Ullrich Melle. Dordrecht, Boston, and London: Kluwer Academic Publishers, 2002, supplementary volume, part one (= Husserliana, vol. XX/1). Imorde, Joseph. Präsenz und Repräsentanz oder: die Kunst, den Leib Christi auszustellen. Das vierzigstündige Gebet von den Anfängen bis in das Pontifikat Innocenz X. Emsdetten and Berlin: Edition Imorde, 1997. Institut de France-Académie des Sciences, ed. Congrès astrophotographique international tenu à l’Observatoire de Paris pour le levé de la Carte du Ciel. Paris: Gauthier-Villars, 1887.
534
Bibliography
Irlam, Shaun. Elations. The Poetics of Enthusiasm in Eighteenth-Century Britain. Stanford: Stanford University Press, 1999. Isler, Hansruedi. Thomas Willis. Ein Wegbereiter der modernen Medizin. Stuttgart: Wissenschaftliche Verlagsgesellschaft, 1965. Israel, Jonathan. The Dutch Republic. Its Rise, Greatness, and Fall 1477-1806. Oxford: Clarendon Press, 1995. Ittershagen, Ulrike. Lady Hamiltons Attitüden. Mainz: Philipp von Zabern, 1985. Jachec, Nancy. The Philosophy and Politics of Abstract Expressionism, 1940-1960. Cambridge: Cambridge University Press, 2000. Jacobi, Carl Gustav Jacob. “Laudatio für Franz Liszt.” Wissenschaftliche Monatsblätter IV (1842): 175-76. Jacobi, Moritz Hermann. “Über die Benutzung der Naturkräfte zu menschlichen Arbeiten.” Vorträge aus dem Gebiete der Naturwissenschaften und der Oekonomie gehalten vor einem Kreise gebildeter Zuhörer in der physikalisch-ökonomischen Gesellschaft zu Königsberg. Ed. Karl Ernst von Baer. Königsberg, 1834. 99-123. Jacobi, Moritz Hermann. “Von der eigentlichen Bedeutung des Luxus und der Mode” (unpublished lecture manuscript). St. Petersburg Academy of Sciences Archive, fond 187/op.1/dele 330. St. Petersburg, undated [ca. 1835]. Jacobus de Voragine. Legenda aurea. Jena: Diedrichs, 1925. Jakob, Karlheinz. Maschine, Mentales Modell, Metapher. Studien zur Semantik und Geschichte der Techniksprache. Tübingen: Niemeyer, 1991. Jay, Martin. The Dialectical Imagination. A History of the Frankfurt School and the Institute of Social Research, 1923-1950. Boston: Little, Brown & Co., 1973. Jenkins, Ian. “‘Contemporary Minds.’ Sir William Hamilton’s Affair with Antiquity.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. Jenkins, Ian. “Talking Stones. Hamilton’s Collection of Carved Gems.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. Joblot, Louis. Observations d’histoire naturelle, faites avec le microscope, sur un grand nombre d’Insectes . . . Paris, 1754. Johnson, George. “Here They Are, Science’s 10 Most Beautiful Experiments.” New York Times (Tuesday, September 24, 2002). Johnson, Samuel. A Dictionary of the English Language. London: Knapton, 1755. Johnson, Steven. “Introduction. A Junction at Eighth Street.” The New York Schools of Music and Visual Art. Ed. idem. New York: Routledge, 2002. 1-15. Johnston, Sarah Iles. “Songs for the Ghosts.” The World of Ancient Magic. Ed. David R. Jordan, Hugo Montgomery, and Einar Thomassen. Bergen: The Norwegian Institute at Athens, 1999. Jones, Caroline A. Machine in the Studio. Constructing the Postwar American Artist. Chicago: University of Chicago Press, 1996. Jorissen, Willem P. Het chemisch (thans anorganisch chemisch) laboratorium der universiteit te Leiden van 1859-1909 en de chemische laboratoria dier universiteit vóór dat tijdvak en hen, die er in doceerden. Leiden: A.W. Sijthoff’s, 1909. Josephson, Paul R. New Atlantis Revisited. Princeton: Princeton University Press, 1997. Kahng, Eik. “L’Affaire Greuze and the Sublime of History Painting.” Art Bulletin 86 (March 2004): 96-114. Kant, Immanuel. “Untersuchung über die Deutlichkeit der Grundsätze der natürlichen Theologie und der Moral.” Moses Mendelssohn. Abhandlung über die Evidenz in metaphysischen Wissenschaften, welche den von d. Kgl. Acad. d. Wiss. in Berlin auf d. Jahr 1763 ausgesetzten Preis erhalten hat. Berlin, 1764.. § 2, 73.
Bibliography
535
Kant, Immanuel. Werkausgabe. 12 vols. Ed. Wilhelm Weischedel. Frankfurt a.M.: Suhrkamp, 1964. Kant, Immanuel. Critique of Pure Reason. Trans. Norman Kemp Smith. New York: St. Martin’s Press, 1965. Kant, Immanuel. Critique of Judgment. Trans. James Creed Meredith. Oxford: Clarendon Press, 1989. Kant, Immanuel. Political Writings. Ed. H.S. Reiss. Trans. H.B. Nisbet. Cambridge: Cambridge University Press, 1991. Kant, Immanuel. Prolegomena to any Future Metaphysics. Ed. Gary Hatfield. Cambridge: Cambridge University Press, 1997. Kanz, Roland. Die Kunst des Capriccio. Berlin: Deutscher Kunstverlag, 2002. Kaplan, Robert. The Nothing that Is. A Natural History of Zero. London: Penguin, 2001. Kemp, Martin. The Science of Art. Optical Themes in Western Art from Brunelleschi to Seurat. New Haven and London: Yale University Press, 1990. Kemp, Martin. Visualizations. The Nature Book of Art and Science. Berkeley: University of California Press, 2001. Kemp, Wolfgang. ‘. . . einen wahrhaft bildenden Zeichenunterricht überall einzuführen.’ Zeichnen und Zeichenunterricht der Laien 1500-1870. Ein Handbuch. Frankfurt a.M.: Syndikat, 1979. Kerényi, Carl. Eleusis and Athens. Archetypal Image of Mother and Daughter (= Bollingen Series, LXV). Trans. Ralph Manheim. New York: Pantheon, 1967. Khaled, Sandrina. “Pikturale Graphismen der Technik, 1569-1870.” Bilder in Prozessen (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1,1). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2003. 64-78. Kingsley, April. The Turning Point. The Abstract Expressionists and the Transformation of American Art. New York: Simon & Schuster, 1992. Kircher, Athanasius. Musurgia universalis sive ars magna consoni et dissoni in X. libros digesta. 2 vols. Rome, 1650. Kircher, Athanasius. Oedipus Aegyptiacus . . . 2 vols. Rome, 1653. Kircher, Athanasius. Ars magna sciendi. In XII libros digesta, qua nova et universali methodo per artificiosum combinationum contextum de omni re proposita plurimis et propre infinitis rationibus disputari, omniumque summaria quaedam cognitio comparari potest. Amsterdam, 1669. Kircher, Athanasius. Ars magna lucis et umbrae in mundo, atque adeó universa natura, vires effectusque uti nova, ita varia novorum reconditiorumque specimi-num exhibitione, ad varios mortalium usus, panduntur. Rome, 1646 and Amsterdam, 1671. Kircher, Athanasius. Arca Noê . . . Amsterdam, 1675. Kircher, Athanasius. Musurgia universalis sive ars magna consoni et dissoni in X. libros digesta. Hildesheim and New York: Olms, 1970 [facsimile of the edition Rome, 1650]. [Kircher, Athanasius/Agathus Carion.] ATHANASII KIRCHERS è Soc: JESU Neue Hallvnd Thon-Kunst/ Oder Mechanische Gehaim-Verbindung der Kunst und Natur . . . Jn unsere Teutsche Mutter-Sprach übersetzet von AGATHO CARIONE. Hanover: “Libri rari” Schäfer, 1984 [facsimile of the edition Nördlingen, 1684]. [Kircher, Athanasius/Andreas Hirsch.] Artis magnae de consono & dißono ARS MINOR; Das ist/ Philosophischer Extract und Auszug/ aus deß Welt-berühmten Teutschen Jesuitens Athanasii Kircheri . . . MUSURGIA UNIVERSALI . . . Ausgezogen und verfertiget . . . von Andrea Hirschen/ . . . Evangel. Pfarrern zu Bächlingen . . . Leipzig: Zentralantiquariat der DDR, 1988 [facsimile of the edition Schwäbisch Hall, 1662]. Klein, Wolf Peter: “Franciscus Mercurius van Helmont.” Die Philosophie des 17. Jahr-
536
Bibliography
hunderts (= Grundriss der Geschichte der Philosophie). Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Vol. 4.1. Basel: Schwabe, 2001. 48-53. Klever, Wim, ed. Die Schwere der Luft in der Diskussion des 17. Jahrhunderts. Wiesbaden: Harrassowitz, 1997. Klingsöhr-Leroy, Cathrin. Das Künstlerbildnis des Grand siècle in Malerei und Graphik. Vom “Noble Peintre” zum “Pictor Doctus.” Munich: Fink, 2002. Knauer, Peter. Ignatius von Loyola. Geistliche Übungen und erläuternde Texte. Graz: Styria, 1978. Knorr-Cetina, Karin D. The Manufacture of Knowledge. Oxford: Pergamon, 1981. Koerner, Joseph Leo. The Moment of Self-Portraiture in German Renaissance Art. Chicago: University of Chicago Press, 1993. Kohler, Robert E. Lords of the Fly. Chicago: University of Chicago Press, 1994. Kohler, Robert E. Landscapes and Labscapes. Chicago: University of Chicago Press, 2002. Konersmann, Ralf. Der Schleier des Timanthes. Perspektiven der historischen Semantik. Frankfurt a.M.: Fischer, 1994. Kopp, Hermann. Die Alchemie in älterer und neuerer Zeit. Ein Beitrag zur Kulturgeschichte. Hildesheim: Olms, 1962 [facsimile of the edition Heidelberg, 1886]. Körner, Hans. Auf der Suche nach der ‘wahren Einheit.’ Ganzheitsvorstellungen in der französischen Malerei und Kunstliteratur vom mittleren 17. bis zum mittleren 19. Jahrhundert. Munich: Fink, 1988. Krafft, Fritz. Dynamische und statische Betrachtungsweise in der antiken Mechanik. Wiesbaden: Steiner, 1970. Krämer, Sybille. Berechenbare Vernunft. Kalkül und Rationalismus im 17. Jahrhundert. Berlin and New York: de Gruyter, 1991. Krämer, Sybille. “Kalküle als Repräsentation. Zur Genese des operativen Symbolismus in der Neuzeit.” Räume des Wissens. Repräsentation, Codierung, Spur. Ed. HansJörg Rheinberger, Michael Hagner, and Bettina Wahrig-Schmidt. Berlin: Akademie Verlag, 1997. 111-22. Krämer, Sybille. “Zentralperspektive, Kalkül, Virtuelle Realität: Sieben Thesen über die Weltbildimplikationen symbolischer Formen.” Medien-Welten, Wirklichkeiten. Ed. Gianni Vattimo and Wolfgang Welsch. Munich: Fink, 1998. 27-37. Krämer, Sybille. “Die Rationalisierung der Visualität und die Visualisierung der Ratio. Zentralperspektive und Kalkül als Kulturtechniken des ‘geistigen Auges.’” Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst. Ed. Helmar Schramm et al. Berlin: Dahlem University Press, 2003. 50-67. Krämer, Sybille. “‘Schriftbildlichkeit’ oder: Über eine (fast) vergessene Dimension der Schrift.” Bild, Schrift, Zahl. Ed. idem and Horst Bredekamp. Munich: Fink, 2003. 157-76. Krauss, Werner, ed. Fontenelle und die Aufklärung. Munich: Fink, 1969 Krementz, Jill. The Writer’s Desk. New York: Random House, 1996. Kristensen, Leif Kahl. “T.N. Thiele and the Carte du Ciel.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 59-63. Krohn, Wolfgang. “Technik, Kunst und Wissenschaft. Die Idee einer konstruktiven Naturwissenschaft des Schönen bei Leon Battista Alberti.” Leonardo da Vinci. Natur im Übergang. Beiträge zu Kunst, Wissenschaft und Technik. Ed. Frank Fehrenbach. Munich: Fink, 2002. 37-56. Krysmanski, Bernd W. Hogarth’s “Enthusiasm Delineated,” Nachahmung als Kritik am Kennertum. Eine Werkanalyse. Zugleich ein Einblick in das satirisch-aufgeklärte
Bibliography
537
Denkens eines “Künstlerrebellen” im englischen 18. Jahrhundert. Hildesheim: Olms, 1996. Kubik, Gerhard. Theory of African Music. 2. vols. Wilhelmshaven: Noetzel, 1994. Kubik, Gerhard. Kalimba, Nsansa, Mbira – Lamellophone in Afrika. Berlin: Museum für Völkerkunde, 1998. Kuckertz, Josef. “Einleitung“. Außereuropäische Musik in Einzeldarstellungen. Ed. Edition MGG. Kassel, Basel, London: Bärenreiter Verlag, 1980. Kugler, Hartmut. “Die Ebsdorfer Weltkarte. Ein europäisches Weltbild im deutschen Mittelalter.” Zeitschrift für deutsches Altertum und deutsche Literatur 116 (1987): 1-29. Kuh, Katharine. The Artist’s Voice. New York: Harper and Row, 1960. Kuklick, Henrika. “Chicago Sociology and Urban Planning Policy.” Theory and Society 9 (1980): 821-45. Kümmel, Werner. “Puls und Musik (16.-18. Jahrhundert).” Medizinhistorisches Journal 3 (1963): 269-93. Kurtz, Lester R. Evaluating Chicago Sociology. Chicago: University of Chicago Press, 1984. Kutzer, Michael. Anatomie des Wahnsinns. Geisteskrankheit im medizinischen Denken der frühen Neuzeit und die Anfänge der pathologischen Anatomie. Hürtgenwald: Guido Pressler, 1998. La Bruyère, Jean de. Œuvres complètes. Ed. Julien Benda. Paris: Gallimard, 1951. La Bruyère, Jean de. Les caractères. Paris, 1976. La Mettrie, Julien Offray de. “L’homme machine.” Œuvres Philosophiques. 2nd ed. Berlin, 1774. Vol. 1, 275-356. Labat, Jean-Bapatiste. Pater Labats Sklavenbericht. Abenteuerliche Jahre in der Karibik 1690-1705. Ed. Heinrich Pleticha. Stuttgart: Edition Erdmann, 1984. Lamberini, Daniela. “Machines in Perspective. Technical Drawings in Unpublished Treatises and Notebooks of the Italian Renaissance.” The Treatise on Perspective. Published and Unpublished (= Studies in the History of Art 59 – Symposium Papers XXXVI). Ed. Lyle Massey. New Haven and London: Yale University Press, 2003. 213-33. Langgärtner, G. and G. May. “Akklamation.” Lexikon des Mittelalters. Munich: Deutscher Taschenbuch Verlag, 2002. Vol. I, 231f. Lankford, John. “Amateurs and Astrophysics. A Neglected Aspect in the Development of a Scientific Specialty.” Social Studies of Science 11 (1981): 275-303. Lankford, John. “The Impact of Photography on Astronomy.” Astrophysics and Twentieth-Century Astronomy to 1950. Ed. Owen Gingerich. Cambridge and New York: Cambridge University Press, 1984. 16-39. Lankford, John. “Photography and the Nineteenth-Century Transits of Venus.” Technology and Culture 28 (1987): 648-57. Läpple, Christian. “Ghanaische Musik im Spiegel der Literatur des 19. und 20. Jahrhunderts.” Weltmusik II. Ed. Peter Ausländer and Johannes Fritsch. Cologne: Feedback Studio Verlag, 1983. 3-18. Latour, Bruno. Science in Action How to Follow Scientists and Engineers through Society. Cambridge: Harvard University Press, 1987. Latour, Bruno. Pandora’s Hope. Cambridge: Harvard University Press, 1999. Laufhütte, Harmut. “Das Friedensfest in Nürnberg 1650.” 1648. Krieg und Frieden in Europa. Ed. Klaus Bußmann and Heinz Schilling. Munich: Bruckmann, 1998. Vol. 2, 347-58. Laurenza, Domenico. “Leonardo. Le macchine volanti.” Le macchine del Rinascimento. Ed. Giovanni Morello. Rome: Retablo, 2000. 145-87.
538
Bibliography
Laurenza, Domenico. “Corpus mobile. Ansätze einer Pathognomik bei Leonardo.” Leonardo da Vinci. Natur im Übergang. Beiträge zu Kunst, Wissenschaft und Technik. Ed. Frank Fehrenbach. Munich: Fink, 2002. 257-301. Laurenza, Domenico. “Leonardo nella Roma di Leone X (c. 1513-16).” XLIII Lettura Vinciana. Florence: Giunti, 2004. Laurenza, Domenico. Le macchine di Leonardo. Florence: Giunti, 2005. Lawrenson, Thomas Edward. The French Stage and the Playhouse in the XVIIth Century. A Study in the Advent of the Italian Order. New York: Ams Press, 1986. Le Clerc, Jean. “Éloge de feu Mr. de Volder, Professeur en Philosophie et aux Mathématiques, dans l’Académie de Leide.” Bibliothèque choisie 18 (1709): 346-401. Lebrun, Pierre. “Recueil des essaies des merveilles de la peinture de Pierre Lebrun, Peintre (Brussels Manuscript 1635).” Medieval and Renaissance Treatises on the Arts of Painting. Original Texts with English Translations. Ed. Mary P. Merrifield. New York: Dover, 1999. 766-841. Leibniz, Gottfried Wilhelm. Untitled fragment [1677]. Die philosophischen Schriften von Gottfried Wilhelm Leibniz. Ed. Carl Immanuel Gerhardt. Berlin: Weidmann, 187590. Vol. 7, 184-89. Leibniz, Gottfried Wilhelm. “Projet de Préface à la science générale.” Die philosophischen Schriften von Gottfried Wilhelm Leibniz. Vol. 7. Ed. Carl Immanuel Gerhardt. Hildesheim: Olms, 1890. Leibniz, Gottfried Wilhelm. Opuscules et fragments inédits de Leibniz. Extraits des manuscrits de la Bibliothèque Royale de Hanovre. Ed. Louis Couturat. Paris: F. Alcan. 1903. Leibniz, Gottfried Wilhelm. “Drôle de Pensée, touchant une nouvelle sorte de REPRESENTATIONS.” Sämtliche Schriften und Briefe. Ed. Akademie der Wissenschaften zu Berlin. Berlin: Akademie Verlag, 1970. Series IV, vol. I, 562-68 and notes 69496. Leibniz, Gottfried Wilhelm. New Essays on Human Understanding. Ed. and trans. Peter Remnant and Jonathan Bennet. Cambridge: Cambridge University Press, 1996. Leibniz, Gottfried Wilhelm. Philosophical Texts. Ed. and trans. R.S. Woolhouse and Richard Francks. Oxford: Oxford University Press, 1998. Leinkauf, Thomas. “Athanasius Kircher.” Die Philosophie des 17. Jahrhunderts. (= Grundriss der Geschichte der Philosophie). Vol. 4.1. Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Basel: Schwabe, 2001. 269-85. Lemoine, Annick. “Nicolas Régnier, zugeschrieben, Porträt des Marquese Vincenzo Giustiniani.” Caravaggio. Die Sammlung Giustiniani und die Berliner Gemäldegalerie in Preussen [Exhib. cat.]. Ed. Silvia Danesi Squarzina, Milan: Electa, 2001. 198-201. Lenoir, Timothy. Instituting Science. The Cultural Production of Scientific Disciplines. Stanford: Stanford University Press, 1997. Lenzen, Victor F. “Charles Peirce as Astronomer.” Studies in the Philosophy of Charles Sanders Peirce. Ed. Edward C. More and Richard S. Robin. Amherst: University of Massachusetts Press, 1964. 33-50. Lenzen, Victor F. “The Contributions of Charles S. Peirce to Metrology.” Proceedings of the American Philosophical Society 109 (1965): 29-46. Lenzen, Victor F. “The Role of Science in the Philosophy of C.S. Peirce.” Logik, Erkenntnis- und Wissenschaftstheorie, Sprachphilosophie, Ontologie und Metaphysik. Akten des XIV. Internationalen Kongresses für Philosophie, 2-9 September 1968. Vienna: Herder, 1969. 371-76. Lenzen, Victor F. “Charles S. Peirce as Mathematical Geodist.” Transactions of the Charles S. Peirce Society 8 (1972): 90-105.
Bibliography
539
Leonardo da Vinci. Il Codice Atlantico di Leonardo da Vinci nella Biblioteca Ambrosiana di Milano. Ed. Accademia dei Lincei. Transcribed by Augusto Marinoni. 24 vols. Florence: Barbèra, 1973-1980. Leonardo da Vinci. Codices Madrid. Ed. Ladislao Reti and Augusto Marinoni. 5 vols. Frankfurt a.M.: Fischer, 1974. Leonardo da Vinci. Libro di Pittura. Ed. Carlo Pedretti. Florence: Giunti, 1995. Leonardo da Vinci. Il codice Arundel 263 nella British Library. Ed. Carlo Pedretti. Transcribed with commentary by Carlo Vecce. Florence: Giunti, 1998. Leroi-Gourhan, André. Gesture and Speech. Trans. Anna Bostock. Cambridge: MIT Press, 1993. Lestel, Dominique. Les origines animales de la culture. Paris: Flammarion, 2001. Letsinger-Miller, Lyn. The Artists of Brown County. Bloomington: Indiana University Press, 1994. Leutmann, Johann Georg. Neue Anmerckungen vom Glaßschleiffen, darinnen die rechten Maschinen die Gläßer durch Hülffe dreyer Bewegungen zu mehrerer Vollkommenheit zubringen . . . Wittenberg, 1719. Lincoln, Bruce. Emerging from the Chrysalis. Studies in Rituals of Women’s Initiations. Cambridge: Harvard University Press, 1981. Lincoln, Bruce. Priests, Warriors, and Cattle. A Study in the Ecology of Religion. Berkeley, Los Angeles, and London: University of California Press, 1981. Lincoln, Bruce. Theorizing Myth. Narrative, Ideology, and Scholarship. Chicago and London: University of Chicago Press, 1999. Lindeboom, Gerrit A. Herman Boerhaave. The Man and his Work. Leiden and London: Methuen, 1968. Lindeboom, Gerrit A. “Dog and Frog. Physiological Experiments at Leiden during the Seventeenth Century.” Leiden University in the Seventeenth Century. An Exchange of Learning. Ed. Theodor H. Lunsigh Scheurleer and Guillaume H.M. Posthumus Meyjes. Leiden: Brill, 1975. 279-93. Lindner, Rolf. The Reportage of Urban Culture. Robert Park and the Chicago School. Cambridge: Cambridge University Press, 1996. Lippe, Rudolf zur. Vom Leib zum Körper. Naturbeherrschung am Menschen in der Renaissance. Reinbek bei Hamburg: Rowohlt, 1988. Lippmann, Edmund Oskar von. Entstehung und Ausbreitung der Alchemie. 3 vols. Berlin: Springer, 1919. Livingstone, David N. Putting Science in Its Place. Chicago: University of Chicago Press, 2003. Luhmann, Niklas. Die Wirtschaft der Gesellschaft. Frankfurt a.M.: Suhrkamp, 1988. Lünig, Johann Christian. Theatrum ceremoniale historico-politicum, oder Historischund Politischer Schau-Platz aller Ceremonien. 2 [3] vols. Leipzig, 1719/20. Lunsingh Scheurleer, Theodor H. “Un amphithéâtre d’anatomie moralisée.” Leiden University in the Seventeenth Century. An Exchange of Learning. Ed. idem and Guillaume H.M. Posthumus Meyjes. Leiden: Brill, 1975. 217-77. Lunsingh Scheurleer, Theodor H., Cornelia W. Fock, and Albert J. van Dissel. Het Rapenburg. Geschiedenis van eene Leidsche gracht. 15 vols. Leiden: Universiteit Leiden, 1986-1992. Lyons, Claire. “The Neapolitan Context of Hamilton’s Antiquities Collection.” Journal of the History of Collections 9 (No. 2, 1997): 229-40. Macagno, Matilde. Geometry in Motion in the Manuscripts of Leonardo da Vinci. Iowa City: University of Iowa Press, 1987. Maffei, Scipione. “Nuova invenzione d’un gravecembalo col piano e forte; aggiunte al-
540
Bibliography
cune considerazioni sopra gli strumenti musicali.” Giornale de’ letterati d’Italia 5 (Venice, 1711): 144-59. Malthête, Jacques and Laurent Mannoni, eds. Méliès. Magie et cinéma. Paris: Éditions Paris-Musées, 2002. Mann, Heinz Herbert. Augenglas und Perspektiv. Studien zur Ikonographie zweier Bildmotive. Berlin: Gebr. Mann, 1992. Mannheim, Karl. Ideology and Utopia. New York: Harcourt Brace Jovanovich, 1936. Mannoni, Laurent. Étienne-Jules Marey. La mémoire de l’œil. Paris and Milan: Mazotta/ La Cinémathèque Française, 1999. Mansfeld, Jaap, ed. Die Vorsokratiker I. Stuttgart: Reclam, 1987. Marani, Pietro C. “Leonardo dalla scienza all’arte; un cambiamento di stile, gli antefatti, una cronologia.” Fra Rinascimento, Manierismo e Realtà. Scritti di storia dell’arte in memoria di Anna Maria Brizio. 2nd ed. Ed. idem. Florence: Barbèra, 1984. 41-52. Marey, Étienne-Jules. La machine animale. Paris, 1873. Marey, Étienne-Jules. La méthode graphique dans les sciences expérimentales et particulièrement en physiologie et en médecine. Paris, 1878. Marey, Étienne-Jules. Le mouvement. Paris, 1894. Martin, Luther H. Hellenistic Religions. An Introduction. New York: Oxford University Press, 1987. Marx, Karl. “A Contribution to the Critique of Political Economy. Part One.” Collected Works. 50 vols. Moscow: Progress Publishers, 1975-2005. Vol. 29 (1987): Economic Works, 1857-61. 257-420. Maschat, Herbert. Leonardo da Vinci und die Technik der Renaissance. Munich: Profil, 1989. Maurice, Klaus. Der drechselnde Souverän. Materialien zu einer fürstlichen Maschinenkunst. Zurich: Ineichen, 1985. Mayer, Andreas. “Autographien des Ganges. Repräsentation und Redressement bewegter Körper im neunzehnten Jahrhundert.” Kunstmaschinen. Spielräume des Sehens zwischen Wissenschaft und Ästhetik. Ed. idem and Alexandre Métraux. Frankfurt a.M.: Fischer, 2005. 101-138. Mayr, Otto. “Automatenlegenden in der Spätrenaissance.” Technikgeschichte 41.1 (1974): 20-32. McGahagan, Thomas A. Cartesianism in the Netherlands, 1639-1676. The New Science and the Counter-Reformation. Ph.D. Pittsburgh, 1976. Meer, John Henry van der. “Instrumentenkunde.” Die Musik in Geschichte und Gegenwart. Ed. Friedrich Blume et al. 2nd rev. ed. Kassel: Bärenreiter, 1996. Vol. 4, 951-70. Meier-Oeser, Stephan. “Roger Bacon oder Doctor mirabilis und die Macht der Wahrheit.” Zwischen Narretei und Weisheit. Biographische Skizzen und Konturen alter Gelehrsamkeit. Ed. Gerald Hartung and Wolf Peter Klein. Hildesheim: Olms, 1997. 95-124. Meier-Oeser, Stephan. “Einleitung zu § 1. Hermetisch-platonische Naturphilosophie.” Die Philosophie des 17. Jahrhunderts. (= Grundriss der Geschichte der Philosophie). Vol. 4.1. Ed. Helmut Holzhey and Wilhelm Schmidt-Biggemann. Basel: Schwabe, 2001. 7-18. Meinel, Christoph. “Die Chemie an den Universitäten des 18. Jahrhunderts – Institutionalisierungsstufen und konzeptioneller Wandel.” Academia Analecta. Mededelingen van de Koninklijke Academie voor Wetenschappen, Letteren en Schone Kunsten van België. Klasse der Wetenschappen 48 (1986): 37-57. Melmed, Allan. “The Art and Science and Other Aspects of Making Sharp Tips.” Journal of Vacuum Science and Technology B 9.2 (1991): 601-08.
Bibliography
541
Mendelson, Elliott. Introduction to Mathematical Logic. Princeton: Van Nostrand, 1964. Menestrier, Claude-François. Traité des Tournois, Ioustes, Caroussels, et autres Spectacles publics. Lyon, 1664. Menestrier, Claude-François. Des décorations funèbres . . . Paris, 1683. Mercure galant. Paris, 1673. Vol. III. Merolla, Girolamo. Breve, e svccinta Relatione del Viaggio nel regno die Congo nell’ Africa Meridionale. Naples, 1687. Mersch, Dieter. “Ästhetischer Augenblick und Gedächtnis der Kunst. Überlegungen zum Verhältnis von Zeit und Bild.” Die Medien der Künste. Beiträge zur Theorie des Darstellens. Ed. Dieter Mersch. Munich: Fink, 2003. 151-76. Mersenne, Marin. Harmonie universelle. The Books on Instruments. Trans. Roger E. Chapman. The Hague: Nijhoff, 1957. Mersenne, Marin. Harmonie universelle. Ed. François Lesure. Paris: Centre national de la recherche scientifique, 1963 [facsimile of the edition Paris, 1636]. Merton, Robert K. Social Theory and Social Structure. New York: Free Press, 1968. Meschkowski, Herbert, ed. Lust an der Erkenntnis. Moderne Mathematik. Munich and Zurich: Piper, 1991. Mesnadière, Jules de la. La Poëtique. Paris, 1640. Meya, Jörg and Heinz Otto Sibum. Das fünfte Element. Wirkungen und Deutungen der Elektrizität. Reinbek b. Hamburg: Rowohlt, 1987. Meyer, Alfred. Historical Aspects of Cerebral Anatomy. New York and London: Oxford University Press, 1971. Meyer, Alfred and Raymond Hierons. “On Thomas Willis’s Concepts of Neurophysiology.” Medical History 9 (1965): 1-15, 142-155. Meyer, Andreas. Afrikanische Trommeln. West- und Zentralafrika. Berlin: Museum für Völkerkunde, 1997. Meyer, Klaus. Geheimnisse des Antoni von Leeuwenhoek. Ein Beitrag zur Frühgeschichte der Mikroskopie. Lengerich: Pabst, 1998. Meyer, Rudolf J. Königs- und Kaiserbegräbnisse im Spätmittelalter. Von Rudolf von Habsburg bis zu Friedrich III. (= Forschungen zur Kaiser- und Papstgeschichte des Mittelalters, vol. 19). Cologne, Weimar, and Vienna: Böhlau, 2000. Meyers Großes Konversations-Lexikon. 6th ed. Leipzig and Vienna, 1908. Meyers Enzyklopädisches Lexikon. 9th ed. Mannheim, Vienna, and Zurich, 1976. Michel, Christian. Le Voyage d’Italie de Charles-Nicolas Cochin. Rome: École française de Rome, 1991 [1758]. Mill, John Stuart. “The Spirit of the Age” [1831]. Collected Works of John Stuart Mill. Ed. John M. Robson et al. Toronto: University of Toronto Press, 1981-91. Vol. 32, 227-316. Miniati, Mara, ed. Museo di Storia della Scienza. Catalogo. [Exhibit. cat.] Florence: Olschki, 1991. Ministère de l’Instruction Publique et des Beaux-Arts. Conférence internationale du passage de Vénus. Procès-Verbaux. Paris: Imprimerie Nationale, 1881. Ministerium für Kultur der Republik Griechenland, ed. Geschenke der Musen. Musik und Tanz im antiken Griechenland. Athens: Ministerium für Kultur der Republik Griechenland, 2003. Mody, Cyrus. “Instruments in Training. The Growth of American Probe Microscopy in the 1980s.” Pedagogy and the Practice of Science. Producing Physical Scientists, 1800-2000. Ed. David Kaiser. Cambridge: MIT, 2005. 185-216. Moffett, Thomas. Insectorvm sive minimorum animalivm theatrum. London, 1634. Molhuysen, Philip C. Bronnen tot de geschiedenis der Leidsche Universiteit. 7 vols. The Hague: Nijhoff, 1913-1924.
542
Bibliography
Molinari, Cesare. Le nozze degli dèi. Un saggio sul grande spettacolo italiano nel seicento. Rome: Bulzoni, 1968. Molland, A. George. “Roger Bacon as Magician.” Traditio 30 (1974): 445-460. Montague, William. The Delights of Holland. Or, A Three Months Travel about that and the other Provinces with Observations and Reflections on their Trade, Wealth, Strength, Beauty, Policy, etc. Together with a Catalogue of the Rarities in the Anatomical School at Leiden. London, 1696. Moran, Bruce T. The Alchemical World of the German Court. Occult Philosophy and Chemical Medicine in the Circle of Moritz of Hessen, 1572-1632. Stuttgart: F. Steiner Verlag, 1991. Morley, Christopher Love. Collectanea chymica Leydensia. Jena, 1696. Morton, Alan and Jane Wess. Public and Private Science. Oxford: Oxford University Press, 1993. Motherwell, Robert et al., eds. Modern Artists in America. First Series. New York: Wittenborn Schultz, 1951. Mouchez, Ernest B. La Photographie astronomique à l’Observatoire de Paris et la Carte du Ciel. Paris: Gauthier-Villars, 1887. Moynet, Jean. L’envers du théatre. Machines et décorations. Béziers: Hachette, 1990 [facsimile of the edition Paris, 1873]. Münster, Sebastian. Fürmalung und künstlich beschreibung der Horologien. Basel, 1537. Murray, Janet H. Hamlet on the Holodeck. The Future of Narrative in Cyberspace. Cambridge and London: MIT Press, 1997. Nagel, Thomas. The View from Nowhere. Oxford: Oxford University Press, 1986. Nagler, Alois. Theatre Festivals of the Medici, 1539-1637. London and New Haven: Da Capo, 1964. Naulin, Paul. “Le problème de la conscience et la notion d’‘image.’” Bergson. Naissance d’une philosophie. Actes du colloque de Clermont-Ferrand. Paris: P.U.F, 1990. 97109. Ndalianis, Angela. Neo-Baroque Aesthetics and Contemporary Entertainment. Cambridge and London: MIT Press, 2004. Needham, Joseph. Science and Civilization in China. 7 vols. Cambridge: Cambridge University Press, 1959ff. Nelle, Florian. “Descartes und der Regenbogen im Wasserglas – Von der beobachteten zur inszenierten Natur.” Theatralität und die Krisen der Repräsentation. Ed. Erika Fischer-Lichte. Stuttgart and Weimar: Metzler, 2001. 392-410. Nelle, Florian. “Im Rausch der Dinge. Poetik des Experiments im 17. Jahrhundert.” Bühnen des Wissens. Interferenzen zwischen Wissenschaft und Kunst. Ed. Helmar Schramm et al. Berlin: Dahlem University Press, 2003. 140-68. Nelson, Benjamin. The Idea of Usury. 2nd expanded edition. Chicago and London: University Press, 1969. Nettl, Bruno. “Ethnomusicology today.” The World of Music 17,4 (1975): 11-15. Neuburger, Max. Historical Development of Experimental Brain and Spinal Cord Physiology before Flourens. Ed. and trans. Edwin Clarke. Baltimore: Johns Hopkins University Press, 1981. Newton, Isaac. Philosophia naturalis principia mathematica. London, 1687. Nicklaus, Hans Georg. “Opern des Barock als technisches Spektakel.” Frühneuzeit-Info 1 (2004): 40-46. Nicolas von Oresme. Traktat über Geldabwertungen. De Mutatione Monetarum: Tractatus. Epilogue by Martin Burckhardt. Berlin: Kadmos, 1999. Nietzsche, Friedrich. Beyond Good and Evil. Trans. R.J. Hollingdale. Intro. Michael Tanner. London: Penguin, 1973.
Bibliography
543
The Nobel Foundation. Press Release: The 1986 Nobel Prize in Physics. Online at: http://www.nobel.se/physics/laureates/1986/press.html (August 2005). Nobis, Heribert M. “Die wissenschaftstheoretische Bedeutung der Quaestiones Mechanicae.” Der Wissenschaftsbegriff. Historische und systematische Untersuchungen. Ed. Alwin Diemer. Meisenheim: Hain, 1970. 47-63. Nollet, Jean Antoine. Rede von der nötigen Geschicklichkeit zur Erforschung der Natur, welche er den 15. Mai 1753 bei dem Antritte seines öffentlichen Lehramtes in dem Navarrischen Collegio gehalten. Erfurt, 1755. Norman, Daniel. “The Development of Astrophotography.” Osiris 5 (1938): 560-94. O’Hora, Nathy P. “Astrographic Catalogues of British Observatories.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 135-38. Oberhuber, Konrad, Jay A. Levenson, and Jacquelyn L. Sheehan. Early Italian Engravings from the National Gallery of Art. Washington: National Gallery of Art, 1973. Oechslin, Werner. “Von Piranesi zu Libeskind. Erklären mit Zeichnung.” Daidalos 1 (1981): 15-19. Oken, Lorenz. Lehrbuch der Naturphilosophie. Zürich: Schultheß, 1843. Olschki, Leo S., ed. Catalogo degli strumenti del museo di storia della scienza. Florence: Olschki, 1954. Orlandi, Pellegrino Antonio. Abecedario pittorico. Venice: Pasquali, 1753. Ossola, Carlo. Dal “Cortegiano” all’ “Uomo di mondo.” Storia di un libro e di un modello sociale. Torino: Einaudi, 1987. Ouvaroff, Serge. [Sergei Semenovich, Count Uvarov (1786-1855)]. Essay on the Mysteries of Eleusis. Trans. J.D. Price with observations by James Christie. London: Rodwell and Martin, 1817. Ovitt, George. The Restoration of Perfection. Labor and Technology in Medieval Culture. New Brunswick and London: Rutgers University Press, 1987. Pagel, Walter. William Harvey’s Biological Ideas. Basel, New York: S. Karger, 1966. Palmer, Vivien M. Field Studies in Sociology. A Student’s Manual. Chicago: University of Chicago Press, 1928. Pancaldi, Giuliano. Volta. Science and Culture in the Age of Enlightenment. Princeton: Princeton University Press, 2003. Paniagua, Cecilio. “Notes on a Drawing by Leonardo da Vinci.” International Review of Psychoanalysis 13 (1986): 445-52. Panofsky, Erwin. “Die Perspektive als ‘symbolische Form.’” Vorträge der Bibliothek Warburg 25 (1924), reprinted in H. Oberer and E. Verheyen, eds. Aufsätze und Grundfragen der Kunstwissenschaft. Berlin: Spiess, 1992. 258-330. Paré, Ambroise. Les œuvres . . . de l’anatomie que des instruments de chirurgie. Paris, 1585. Park, Robert E. and Ernest W. Burgess. The City. Chicago: University of Chicago Press, 1925. Park, Sang-il, Jun Nogami, and Cal F. Quate. “Effect of Tip Morphology on Images Obtained by Scanning Tunneling Microscopy.” Physical Review B 36.5 (1987): 2863-66. Parmentier, Marc, ed. L’estime des apparences. 21 manuscripts de Leibniz sur les probabilités, la théorie des jeux, l’espérance de vie. Paris: Vrin, 1995. Partington, James Riddick. A History of Chemistry. Vol. 2. London: Macmillan, 1961. Pastor, Ludwig von, ed. Antonio de Beatis. Die Reise des Kardinals Luigi d’Aragona durch Deutschland, die Niederlande, Frankreich und Oberitalien 1517 bis 1518. Freiburg: Herdersche Verlagshandlung, 1905. Paulhan, Jean. La peinture cubiste (1942-1956). Paris: Denoël, 1990.
544
Bibliography
Pedretti, Carlo. Leonardo da Vinci on Painting. A Lost Book (Libro A). Berkeley and Los Angeles: University of California Press, 1964. Pedretti, Carlo. The Literary Works of Leonardo da Vinci. Commentary. 2 vols. London: Phaidon, 1977. Peirce, Charles S. “On the Proper Functions of a National Office of Weights and Measures” [1899]. Unpublished MS., MSS. Charles S. Peirce, Houghton Library, Harvard University. Peirce, Charles S. Collected Papers. Vol. 5. Cambridge: Harvard University Press, 1934. Peirce, Charles S. MS 204 [Fall 1872]. “Chapter IV of Reality.” Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 54-59. Peirce, Charles S. “Consequences of Four Incapacities” [1868]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 2, 239. Peirce, Charles S. “General Remarks upon Gravity Determinations, by John Herschel.” Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 365-68. Peirce, Charles S. “Grounds of Validity” [1869]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 2, 242-72. Peirce, Charles S. “Measurements of Gravity at Initial Stations in America and Europe.” [1879]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 79-144. Peirce, Charles S. “Six Reasons for the Prosecution of Pendulum Experiments” [1882]. Writings of Charles S. Peirce. A Chronological Edition. Ed. Christian J.W. Kloesel et al. Bloomington: Indiana University Press, 1986. Vol. 4, 359-60. Perrault, Claude. Mémoires pour servir à l’histoire naturelle des animaux. Paris, 1671. Perrault, Claude. Essais de physique. Vol. 3. Paris, 1688. Perrault, Charles. Les hommes illustres qui ont paru en France pendant ce siècle . . . Paris: Dezallier, 1700. Persons, Stow. Ethnic Studies at Chicago, 1905-45. Urbana: University of Illinois Press, 1987. Petri, Richard Julius. Das Mikroskop. Von seinen Anfängen bis zur jetzigen Vervollkommnung für alle Freunde dieses Instruments. Berlin: Schoetz, 1896. Pfeifer, Rold and Christian Scheter. Understanding Intelligence. Cambridge and London: MIT Press, 1999. Pfisterer, Ulrich, ed. Die Kunstliteratur der italienischen Renaissance. Eine Geschichte in Quellen. Stuttgart: Reclam, 2002. Pickering, Andrew. The Mangle of Practice. Time, Agency, and Science. Chicago: University of Chicago Press, 1995. Picon, Antoine. French Architects and Engineers in the Age of Enlightenment. Trans. Martin Thom. Cambridge: Cambridge University Press, 1992. Piemme, Jean-Marie. “L’utile dulci ou la convergence des nécessités. Recherches historiques sur les causes de l’adoption de la règle scaligérienne de l’utilité, par les théoriciens de 1630.” Revue d’histoire du théâtre 2 (1969): 118-33. Piérobon, Franck. Kant et les mathématiques. Paris: Vrin, 2003. Pilipczuk, Alexander. Elfenbeinhörner im sakralen Königtum Schwarzafrikas. Bonn: Verlag für Systematische Musikwissenschaft, 1985. Pino, Paolo. Dialogo di pittura di Messer Paolo Pino [1548]. Ed. Susanna Falabella. Rome: Lithos editrice, 2000.
Bibliography
545
Plato. Phaedo. Trans. David Gallop. Oxford: Clarendon Press, 1975. Plato. Republic. Indianapolis: Hackett, 1992. Plato. Phaedrus. Indianapolis: Hackett, 1995. Pliny the Elder, Gaius. Historia naturalis/Naturkunde. Latin/German. Ed. and trans. Roderich König. Munich and Darmstadt: Artemis & Winkler, 1973ff. Pliny the Elder, Gaius. Natural History. Trans. H. Rackham. Cambridge: Harvard University Press, 1979. Poliander. Anmuthige und Zeit-kürtzende aus vielen curieusen und nützlichen Wissenschaften hergeleitete Quellen. Erster - Fünfzehnter Gang. Erfurt, 1720. Pollack, Peter. The Picture History of Photography. New York: Henry N. Abrams, 1977. Popp, Anny E. Leonardo da Vinci. Zeichnungen. Munich: Piper, 1928. Popplow, Marcus. “Die Verwendung von lat. ‘machina’ im Mittelalter und in der Frühen Neuzeit. Vom Baugerüst zu Zoncas mechanischem Bratenwender.” Technikgeschichte 30.1 (1993): 7-26. Popplow, Marcus. Neu, nützlich und erfindungsreich. Die Idealisierung von Technik in der Frühen Neuzeit. Münster: Waxmann, 1998. Popplow, Marcus. “Maschinenzeichnungen der Ingenieure der Renaissance.” Frühneuzeit-Info 1 (2004): 13-32. Power, Henry. Experimental Philosophy. London, 1664. Praetorius, Michael. Syntagma Musicum. De Organographia. 2 vols. Wolfenbüttel, 1619 and 1620. Praz, Mario. Studies in Seventeenth-Century Imagery. Vol. 1. London: Warburg Institute, 1939-1947. Pure, Michel de. Idée des spectacles anciens et nouveaux. Paris, [1668]. Puttfarken, Thomas. The Discovery of Pictorial Composition. Theories of Visual Order in Painting 1400-1800. New Haven and London: Yale University Press, 2000. Quoika, Rudolf. Altösterreichische Hornwerke. Berlin: Merseburger, 1959. Radau, Rudolphe. “Sur les erreurs personnelles.” Moniteur Scientifique-Quesneville (15 November 1865): 22. “Radiobiology in the Atomic Age: Changing Research Practices and Policies in Comparative Perspective”. Journal of the History of Biology 39.4 (2006). Rádl, Emanuel. Geschichte der Biologischen Theorien in der Neuzeit. 2 vols. Leipzig, 1909. Rasmussen, Nicolas. Picture Control. Stanford: Stanford University Press, 1997. Read, John. Prelude to Chemistry. An Outline of Alchemy, its Literature and Relationships. London: G. Bell and Sons, Ltd., 1939. Reeves, Eileen. “Old Wives’ Tales and the New World System. Gilbert, Galileo, and Kepler.” Configurations 7 (Fall 1999): 301-54. Reid, Constance. From Zero to Infinity. What Makes Numbers Interesting. San Francisco: Mathematical Association of America, 1992. Reinhardt, Rüdiger. Wissen als Ressource. Theoretische Grundlagen, Methoden und Instrumente zur Erfassung von Wissen. Frankfurt a.M.: Peter Lang Verlag, 2002. Restle, Conny. Bartolomeo Cristofori und die Erfindung des Hammerclaviers. Munich: Editio Maris, 1990. Restle, Conny. “Cristofori, Bartolomeo.” Die Musik in Geschichte und Gegenwart. Ed. Friedrich Blume et al. 2nd rev. ed. Kassel: Editio Maris, 2001. Vol. 5, 97-103. Restle, Conny. “Idee und Realisierung. Bartolomeo Cristoforis Weg zum Hammerflügel.” Jahrbuch des Staatlichen Instituts für Musikforschung (2004): 110-22. Reti, Ladislao. “Leonardo the Technologist. The Problem of the Prime Mover.” Idem and Bern Dibner. Leonardo da Vinci, Technologist. Norwalk: Burndy Library, 1969.
546
Bibliography
Reynolds, Sir Joshua. Discourses on Art. Ed. Robert W. Wark. San Marino: The Huntington Library, 1959. Rheinberger, Hans-Jörg. “Vom Mikrosom zum Ribosom. ‘Strategien’ der ‘Repräsentation’ 1935-1955.” Die Experimentalisierung des Lebens. Ed. idem and Michael Hagner. Berlin: Akademie Verlag, 1993. 162-87. Rheinberger, Hans-Jörg. “Putting Isotopes to Work. Liquid Scintillation Counters, 19501970.” Instrumentation Between Science, State, and Industry (= Sociology of the Sciences Yearbook). Ed. Bernward Joerges and Terry Shinn. Dordrecht: Kluwer, 2001. 143-74. Rheinberger, Hans-Jörg. “Präparate – ‘Bilder’ ihrer selbst. Eine bildtheoretische Glosse.” Oberflächen der Theorie (= Bildwelten des Wissens. Kunsthistorisches Jahrbuch für Bildkritik, vol. 1,2). Ed. Horst Bredekamp and Gabriele Werner. Berlin: Akademie Verlag, 2003. 9-19. Ridolfi, Carlo. “Vita di Tintoretto.” Le Meraviglie dell’Arte ovvero le vite degli illustri pittori veneti e dello stato. Vol. 2. Padua: Arnaldo Forni Editore, 2002 [1648]. Robinson, David. The Lantern Image. Iconography of the Magic Lantern 1420-1880. East Sussex: Magic Lantern Society, 1993. Rohlfs, Gerhard. Quer durch Afrika. Die Erstdurchquerung der Sahara vom Mittelmeer zum Golf von Guinea 1865-1867. Ed. Herbert Gussenbauer. Stuttgart: Edition Erdmann, 1984. Rohr, Julius Bernhard von. Einleitung zur Ceremoniel-Wissenschafft Der großen Herren. Ed. Monika Schlechte. Leipzig, 1990 [facsimile of the edition Berlin, 1733]. Rohr, Julius Bernhard von. Einleitung zur Ceremoniel-Wissenschafft Der Privat-Personen. Ed. Gotthardt Frühsorge. Leipzig, 1990 [facsimile of the edition Berlin, 1728]. Romeo, Carmelo. “Matérialisme et Déterminisme dans le Traité de la Liberté de Fontenelle.” Le Matérialisme du XVIIIe siècle et la littérature clandestine. Ed. Olivier Bloch. Paris: Vrin, 1982. Rorty, Richard. “Non-reductive Physicalism.” Objectivity, Relativism, and Truth. Philosophical Papers. Vol. 1. Cambridge: Cambridge University Press, 1990. Rosand, David. Drawing Acts. Studies in Graphic Expression and Representation. Cambridge: Cambridge University Press, 2002. Rose, Paul Lawrence and Stillman Drake. “The Pseudo-Aristotelian ‘Questions of Mechanics’ in Renaissance culture.” Studies in the Renaissance 18 (1971): 65-104. Rosenberg, Bernard and Norris Fliegel. The Vanguard Artist. Chicago: Quadrangle Books, 1965. Rosenberg, Harold. “Gottlieb.” Adolph Gottlieb. Paintings 1950-1971. London: Marlborough Gallery, 1971. Rosenberg, Harold. “Tenth Street. A Geography of Modern Art.” Discovering the Present. Chicago: University of Chicago Press, 1973. 100-09. Rosenberg, Harold. Willem deKooning. New York: Harry N. Abrams, 1973. Rossi, Massimiliano. “La peinture guerrière. Artistes et paladins à Venise au XVIIe siècle.” La Jérusalem délivrée du Tasse. Poésie, peinture, musique, ballet. Ed. Giovanni Careri. Paris: Klincksieck, 1999. 67-108. Rossi, Paolo. Clavis universalis. Arti della memoria e logica combinatoria da Lullo a Leibniz. Bologna: Il Mulino, 1983. Roth, Gerhard. Das Gehirn und seine Wirklichkeit. Frankfurt a.M.: Suhrkamp, 1994. Rotman, Brian. Signifying Nothing. Basingstoke: McMillan, 1987. Rubinstein, Nicolai. “Machiavelli and the Decoration of the Hall of the Great Council in the Palazzo Vecchio.” Musagetes. Festschrift für Wolfram Prinz. Ed. Ronald G. Kecks. Berlin: Mann, 1991. 275-85.
Bibliography
547
Rück, Peter. “Die Dynamik mittelalterlicher Zeitmaße und die mechanische Uhr.” Die Mechanik in den Künsten. Studien zur ästhetischen Bedeutung von Naturwissenschaft und Technologie. Ed. Hanno Möbius and Jörg Jochen Berns. Marburg: Jonas Verlag, 1990. 17-30. Rudolph, Gerhard. “Hallers Lehre von der Irritabilität und Sensibilität.” Von Boerhaave bis Berger. Die Entwicklung der kontinentalen Physiologie im 18. und 19. Jahrhundert. Ed. Karl Eduard Rothschuh. Stuttgart: G. Fischer, 1964. Ruestow, Edward G. The Microscope in the Dutch Republic. The Shaping of Discovery. Cambridge: Cambridge University Press, 1996. Rupp, Jan. “Matters of Life and Death. The Social and Cultural Conditions of the Rise of Anatomical Theatres, with Special Reference to Seventeenth-Century Holland.” History of Science 28 (1990): 263-87. Sabbattini, Nicola. Pratica di fabricar scene, e machine ne’ teatri . . . Ravenna, 1638. Santos, Frei João dos. Ethiopia Orienta [1609]. Lisbon: 1891. Saslow, James M. The Medici Wedding of 1589. Florentine Festival as Theatrum Mundi. New Haven and London: Yale University Press, 1996. Sassen, Saskia. The Global City. 2nd ed. Princeton: Princeton University Press, 2001. Sawin, Martica. Surrealism in Exile and the Beginning of the New York School. Cambridge: MIT Press, 1995. Scannelli, Francesco. Il Microcosmo della Pittura. Bologna: Nuova Alfa Editore, 1989 [facsimile of the edition Cesena, 1657]. Scaramuccia, Luigi. Le finezze de’ pennelli italiani, ammirate, e studiate da Girupeno . . . Pavia, 1674. Schaffer, Simon. “Late Victorian Metrology and Its Instrumentation. A Manufactory of Ohms.” Invisible Connections. Instruments, Institutions, and Science. Ed. Robert Bud and Susan E. Cozzens. Washington: Spie Optical Engineering Press, 1992. 2356. Schaffer, Simon. “Self Evidence.” Critical Inquiry 18 (1992): 327-62. Schaffer, Simon. “Metrology, Metrification and Victorian Values.” Victorian Science in Context. Ed. Bernard Lightman. Chicago and London: University of Chicago Press, 1997. 438-74. Schaffer, Simon. “Fish and Ships. Models in the Age of Reason.” Models. The Third Dimension of Science. Ed. Soraya de Chadarevian and Nick Hopwood. Stanford: Stanford University Press, 2004. 71-105. Schapiro, Meyer. “On Some Problems in the Semiotics of Visual Art. Field and Vehicle in Image-Signs.” Theory and Philosophy of Art. Style, Artist, and Society. New York: Braziller, 1994. 1-32. Scheiner, Julius. Die Photographie der Gestirne. Leipzig: Wilhelm Engelmann, 1897. Scheuchzer, Johann Jacob. Kupfer-Bibel, In welcher Die Physica Sacra, Oder Geheiligte Natur-Wissenschafft Derer In Heil. Schrifft vorkommenden Natürlichen Sachen, Deutlich erklärt und bewährt . . . Augsburg and Ulm, 1731-1735. Schickore, Jutta. “(Ab)using the Past for Present Purposes. Exposing Contextual and Trans-Contextual Features of Error.” Perspectives on Science 10 (2002): 433-56. Schickore, Jutta. “Fixierung mikroskopischer Beobachtungen. Zeichnung, Dauerpräparat, Mikrofotografie.” Ordnungen der Sichtbarkeit. Fotografie in Wissenschaft, Kunst und Technologie. Ed. Peter Geimer. Frankfurt a.M.: Suhrkamp, 2002. 285-310. Schiebinger, Londa. The Mind Has No Sex? Women in the Origins of Modern Science. Cambridge and London: Harvard University Press, 1989. Schimank, Hans. “Zur Geschichte der Physik an der Universität Göttingen vor Wilhelm Weber (1734-1830).” Rete. Strukturgeschichte der Naturwissenschaften 2 (1974): 207-52.
548
Bibliography
Schleiden, Matthias Jacob. Die Botanik als inductive Wissenschaft behandelt. Zweite, ganzlich umgearbeitete Auflage der Grundzüge der Wissenschaftlichen Botanik. Zweiter Theil. Morphologie, Organologie. Leipzig: Engelmann, 1846. Schlosser, Julius von. Die Kunst- und Wunderkammern der Spätrenaissance. Ein Beitrag zur Geschichte des Sammelwesens. Leipzig: Klinkhardt und Biermann, 1908. Schmidt-Biggemann, Wilhelm. “Maschine.” Historisches Wörterbuch der Philosophie. Ed. Joachim Ritter. Darmstadt: Wissenschaftliche Buchgesellschaft, 1980. Vol. 5, 790-802. Schmidt-Biggemann, Wilhelm. Topica universalis. Eine Modellgeschichte humanistischer und barocker Wissenschaft. Hamburg: Meiner, 1983. Schmidt-Biggemann, Wilhelm. “Robert Fludds Theatrum memoriae.” Ars memorativa. Zur kulturgeschichtlichen Bedeutung der Gedächtniskunst 1400-1750. Ed. Jörg Jochen Berns and Wolfgang Neuber. Tübingen: Niemeyer, 1993. 154-70. Schmidt-Biggemann, Wilhelm. Philosophia perennis. Historische Umrisse abendländischer Spiritualität in Antike, Mittelalter und Früher Neuzeit. Frankfurt a.M.: Suhrkamp, 1998. Schmitz, Emil-Heinz. Das Mikroskop (= Handbuch zur Geschichte der Optik, Ergänzungsbd. II). Bonn: Wayenborgh, 1989. Schneider, Irmela. “Von der Vielsprachigkeit zur ‘Kunst der Hybridation.’ Diskurse des Hybriden.” Hybridkultur. Medien, Netze, Künste. Ed. idem and Christian W. Thomsen. Cologne: Wienand, 1997. 13-66. Schott, Caspar. Magia Optica. Das ist Geheime und doch naturmässige Gesichts- und Augenlehr. Würzburg, 1671. Schramm, Helmar. Karneval des Denkens. Theatralität im Spiegel philosophischer Texte des 16. und 17. Jahrhunderts. Berlin: Akademie Verlag, 1996. Schulze, Hendrik. “Afrikaner auf der europäischen Opernbühne des 17. und 18. Jahrhunderts.” A Global View of Mozart. Ed. Generalsekretariat “Mozart 2006 Salzburg.” Salzburg, 2002. 8-11. Schumpeter, Joseph A. Das Wesen des Geldes. Göttingen: Vandenhoeck & Ruprecht, 1970. Schütt, Hans-Werner. Auf der Suche nach dem Stein der Weisen. Die Geschichte der Alchemie. Munich: C.H. Beck, 2000. Schütte, Ulrich. “Architekt und Ingenieur.” Architekt und Ingenieur. Baumeister in Krieg und Frieden. [Exhibit. cat.] Wolfenbüttel: Herzog August Bibliothek, 1984. 18-31. Scolari, Massimo. “Elementi per una storia dell’assonometria.” Casabella 550 (March 1984): 42-49. Searle, John R. The Construction of Social Reality. New York: Free Press, 1995. Seidner, David. Artists at Work. Inside the Studios of Today’s Most Celebrated Artists. New York: Rizzoli, 1999. Seife, Charles. Zero. The Biography of a Dangerous Idea. New York: Viking, 2000. Seifert, Herbert. Die Oper am Wiener Kaiserhof im 17. Jahrhundert (= Wiener Veröffentlichungen zur Musikgeschichte, vol. 25). Tutzing: Schneider, 1985. Seitz, William C. Abstract Expressionist Painting in America. Cambridge: Harvard University Press, 1983. Sellers, David. The Transit of Venus. The Quest to Find the True Distance to the Sun. Leeds: Maga Velda Press, 2001. Seng, Nikolaus. Selbstbiographie des P. Athanasius Kircher aus der Gesellschaft Jesu. Fulda: Druck und Verlag der Fuldaer Actiendruckerei, 1901. Seuphor, Michel. “Paris New York 1950.” Modern Artists in America. First Series. Ed. Robert Motherwell et al. New York: Wittenborn Schultz, 1951. 118-22.
Bibliography
549
Shapin, Steven. “The House of Experiment in Seventeenth-Century England.” Isis 79 (1988): 373-404. Shapin, Steven and Simon Schaffer. Leviathan and the Air-Pump. Hobbes, Boyle, and the Experimental Life. Princeton: Princeton University Press, 1985. Shaw, Clifford R. The Jack-Roller. A Delinquent Boy’s Own Story. Chicago: University of Chicago Press, 1930. Shergold, N.D. A History of the Spanish Stage from Medieval Times until the End of the Seventeenth Century. Oxford: Clarendon Press, 1967. Sibum, H. Otto. “Charles-Augustin Coulomb (1736-1806).” Die Großen Physiker. Von Aristoteles bis Kelvin. Ed. Karl von Meyenn. Munich: Beck, 1997. 243-62. Sibum, H. Otto. “Experimentalists in the Republic of Letters.” Science in Context 16 (2003): 89-120. Simic, Charles. “Difference in Similarity.” New York Review (11 March 2004): 21-23. Singer, Wolf. Ein neues Menschenbild? Gespräche über Hirnforschung. Frankfurt a.M.: Suhrkamp, 2003. Small, Jocelyn Penny. “Time in Space. Narrative in Classical Art.” Art Bulletin 81 (December 1999): 563. Smith, Pamela H. “Chemistry and Commerce. Johann Joachim Becher and the Court of the Elector in Munich.” Johann Joachim Becher (1635-1682). Ed. Gotthardt Frühsorge and Gerhard F. Strasser. Wiesbaden: Harrassowitz, 1993. 143-58. Smith, T.V. and Leonard D. White, eds. Chicago. An Experiment in Social Science Research. Chicago: University of Chicago Press, 1929. Snelders, Henricus A.M. “Johann Joachim Becher und sein Gold-aus-Sand-Projekt.” Johann Joachim Becher (1635-1682). Ed. Gotthardt Frühsorge and Gerhard F. Strasser. Wiesbaden: Harrassowitz, 1993. 103-14. Soemmering, Samuel Thomas. Über das Organ der Seele. Amsterdam, 1966 [facsimile of the Königsberg edition, 1796]. Sohm, Philip. Style in the Art Theory of Early Modern Italy. Cambridge: Cambridge University Press, 2001. Sontag, Susan. The Volcano Lover. A Romance. New York: Farrar, Straus, Giroux, 1992. Speck, Josef, ed. Grundprobleme der großen Philosophen. Philosophie des Altertums und des Mittelalters. 5th ed. Göttingen: Vandenhoeck & Ruprecht, 2001. Spivey, Nigel. Enduring Creation. Art, Pain, and Fortitude. Berkeley, Los Angeles, and London: University of California Press, 2001. Stafford, Barbara Maria.“‘Fantastic’ Images. From Unenlightening to Enlightening Images Meant To Be Seen in the Dark.” Aesthetic Illusion. Theoretical and Historical Approaches. Ed. Frederick Burwick and Walter Pape. Berlin and New York: de Gruyter, 1990. 158-79. Stafford, Barbara Maria. Artful Science. Enlightenment Entertainment and the Eclipse of Visual Information. Cambridge and London: MIT Press, 1996. Stafford, Barbara Maria. Visual Analogy. Consciousness as the Art of Connecting. Cambridge and London: MIT Press, 1999. Stafford, Barbara Maria. “Levelling the New Old Transcendence. Cognitive Coherence in the Age of Beyondness.” New Literary History 35.2 (2004): 321-28. Stagl, Justin. A History of Curiosity. The Theory of Travel, 1550-1800. Chur: Harwood, 1995. Stahl, Georg Ernst. Bedencken, Erinnerung und Erlaeuterung über Doctor J. Bechers Natur-Kuendigung der Metallen. Frankfurt and Leipzig, 1723. Stahl, Georg Ernst. Fundamenta Chymica Dogmatico Rationalis et Experimentalis. Nuremberg, 1732.
550
Bibliography
Staley, Richard. “Michelson’s Interferometer. Experiment or Instrument?” Instrument – Experiment. Historische Studien. Ed. Christoph Meinel. Berlin and Diepholz: Verlag für Geschichte der Naturwissenschaft und der Technik, 2000. 192-200. Standage, Tom. The Victorian Internet. The Remarkable Story of the Telegraph and the Nineteenth-Century’s On-Line Pioneers. New York: Walker, 1998. Steiner, Reinhard. “Malerei als Spekulation.” Digitaler Schein. Ästhetik der elektronischen Medien. Ed. Florian Rötzer. Frankfurt a.M.: Suhrkamp, 1991. 435-54. Steinle, Friedrich. “Exploratives vs. theoriebestimmtes Experimentieren: Amperes erste Arbeiten zum Elektromagnetismus.” Experimental Essays – Versuche zum Experiment. Ed. Michael Heidelberger and Friedrich Steinle. Baden-Baden: Nomos, 1988. 272-97. Stensen, Niels. Sur L’anatomie du cerveau à Messieurs de l’Assemblée, qui se fait chez Monsieur Thevenot. Paris, 1669. Stiegler, Bernard. Technics and Time, 1: The Fault of Epimethus. Trans. Richard Beardsworth and George Collins. Stanford: Stanford University Press, 1998. Stieve, Gottfried. Europäisches Hoff-Ceremoniel. Leipzig: Gleditsch, 1715. Stöcklein, Ansgar. Leitbilder der Technik. Biblische Tradition und technischer Fortschritt. Munich: Moos, 1969. Stoll, Erich. “Information- and Image-processing of Scanning-tunneling-microscope Data.” SPIE 599 (1985): 442-450. Stoll, Erich. “Why do ‘Dirty’ Tips Produce Higher-Resolution Images when Graphite is Scanned in a Scannning Tunneling Microscope?” Journal of Physics C – Solid State Physics 21.26 (1988): L921-L924. Strasser, Bruno. La fabrique d’une nouvelle science. La biologie moléculaire à l’âge atomique (1945-1964). Florence: Olschki, 2006. Ströker, Elisabeth. Denkwege der Chemie. Elemente ihrer Wissenschaftstheorie. Freiburg and Munich: Alber, 1967. Ströker, Elisabeth. Theoriewandel in der Wissenschaftsgeschichte. Chemie im 18. Jahrhundert. Frankfurt a.M.: Vittorio Klostermann, 1982. Strong, Donald S. Leonardo on the Eye. An English Translation and Critical Commentary of MS. D in the Bibliothèque Nationale, Paris, with Studies on Leonardo’s Methodology and Theories of Optics. New York: Garland, 1979. Strong, Roy. The Renaissance Garden in England. London: Thames and Hudson, 1998. Summers, David. The Judgment of Sense. Renaissance Naturalism and the Rise of Aesthetics. 2nd ed. Cambridge: Cambridge University Press, 1990. Tawney, R.T. “The Rise of the Gentry.” History and Society. London, Henly, and Boston: Routledge & Kegan Paul, 1978. 85-128. Teresi, Dick. “Zero.” The Atlantic Monthly 280/1 (1997): 88-94. Thackray, John. “The Modern Pliny. Hamilton and Vesuvius.” Vases & Volcanoes. Sir William Hamilton and His Collection. London: British Museum Press, 1996. 65-74. Thijssen-Schoutte, C. Louise. Nederlands Cartesianisme. Amsterdam: N.V. Noord-Hollandsche Uitgevers, 1954. Thistlewood, David. “Historicity and Mythology in American Abstract Expressionism.” American Abstract Expressionism. Ed. idem. Liverpool: Liverpool University Press, 1993. 1-39. Thompson, Charles John Samuel. The Lure and Romance of Alchemy. London: Harrap, 1932. Thorndike, Lynn. A History of Magic and Experimental Science. 8 Vols. New York: Columbia University Press, 1964. Thrasher, Frederic M. The Gang. A Study of 1313 Gangs in Chicago. Chicago: University of Chicago Press, 1927.
Bibliography
551
Todd, Edwin M. The Neuroanatomy of Leonardo da Vinci. Santa Barbara: Capra Press, 1983. Tozer, E.T. “Discovery of an Ammonoid Specimen describes by Robert Hooke.” Notes and Records of the Royal Society of London 44.1 (1990): 3-12. Tumbleson, Raymond D. Catholicism in the English Protestant Imagination. Nationalism, Religion, and Literature, 1660-1745. Cambridge: Cambridge University Press, 1998. Turing, Alan M. “On Computable Numbers, with an Application to the ‘Entscheidungsproblem.’” Proceedings of the London Mathematical Society 2.42 (1937): 230-65 and 2.43 (1937): 544-46. Turner, Anthony John. Of Time and Measurement. Studies in the History of Horology and Fine Technology. Aldershot: Variorum, 1993. Turner, Gerard L’E. Essays on the History of the Microscope. Oxford: Senecio Publications, 1980. Turner, Gerard L’E. Mikroskope. Munich: Callwey Verlag, 1981. Turner, Herbert Hall. The Great Star Map. New York: E.P. Dutton, 1912. U.S. Transit of Venus Commission. Instructions for Observing the Transit of Venus, December 6, 1882. Washington, D.C.: Government Printing Office, 1882. Vasari, Giorgio. Le Vite de’ più eccellenti pittori, scultori e architettori. 6 vols. Ed. Rosanna Bettarini and Paola Barocchi. Florence: Sansoni, 1966-1987. Vasari, Giorgio. The Lives of the Artists. Trans. Julia Conway Bondanella and Peter Bondanella. Oxford: Oxford University Press, 1998. Velasco, Don Antonio Palomino. Las vidas de los pintores y estatuarios eminentes españoles . . . London, 1724. Veldman, Ilja M. and Ger Luijten. Maarten van Heemskerck. (= The New Hollstein. Dutch & Flemish Etchings, Engravings, and Woodcuts 1450-1700). 2 vols. Amsterdam: Sound & Vision Interactive, 1994. Veranzio, Fausto. Machinae Novae Fausti Verantii siceni cum declaratione Latina, Italica, Hispanica, Gallica, et Germanica. Venice, [1600]. Verbeek, Theo. Descartes and the Dutch. Early Reactions to Cartesian Philosophy, 1637-1650. Carbondale and Edwardsville: Southern Illinois University Press, 1992. Verhandlungen der vom 20. bis 29. September 1875 in Paris vereinigten Permanenten Commission der Europäische Gradmessun. Berlin: Reimer, 1875. Vérin, Hélène. La gloire des ingénieurs. L’intelligence technique du XVIe au XVIIIe siècle. Paris: Albin Michel, 1993. Vernant, Jean-Pierre. Myth and Thought among the Greeks. London, Boston, Melbourne, and Henley: Routledge & Kegan Paul, 1983. Vickers, Michael. “Hamilton, Geology, Stone Vases, and Taste.” Journal of the History of Collections 9.2 (1997): 270. Vieussens, Raymond de. Neurographia universalis. Lyon, 1684. Villalpando, Juan Battista. Explanationes in Ezechielis et apparatus urbis ac templi Hierosolymitani. Rome, 1604. Virdung, Sebastian. Musica getutscht. Basel, 1511. Virtual Laboratory for Physiology. Online: http://vlp.mpiwg-berlin.mpg.de Vitruvius. On Architecture. 2 vols. Trans. F. Granger. Cambridge: Harvard Univerisity Press, 1962. Vitruvius. The Ten Books on Architecture. Trans. Ingrid D. Rowland. Cambridge: Cambridge University Press, 1999. Völter, Ulrich. Geschichte und Bedeutung der internationalen Erdmessung. Munich: Verlag der Bayerischen Akademie der Wissenschaften, 1963. Walsh, P.G. The Roman Novel. The “Satyricon” of Petronius and the “Metamorphoses” of Apuleius. Cambridge: Cambridge University Press, 1970.
552
Bibliography
Wanger, Bernd Herbert. Kaiserwahl und Krönung im Frankfurt des 17. Jahrhunderts. Darstellung anhand der zeitgenössischen Bild- und Schriftquellen und unter besonderer Berücksichtigung der Erhebung des Jahres 1612 (= Studien zur Frankfurter Geschichte, vol. 34). Frankfurt a.M.: Kramer, 1994. Warburg, Aby. “I costumi teatrali per gli Intermezzi del 1589. I disegni di Bernardo Buontalenti e il ‘Libro di conti’ di Emilio de’ Cavalieri.” Gesammelte Schriften. Leipzig: Teubner, 1932. 259-300 and 394-441. Warnke, Martin. The Court Artist. On the Ancestry of the Modern Artist. Cambridge: Cambridge University Press, 1993. Warren, Rosanna. Departure. New York: Norton, 2003. Wasson, R. Gordon. “The Divine Mushroom of Immortality.”The Flesh of the Gods. The Ritual Use of Hallucinogens. Ed. P.T. Furst. New York: Praeger, 1972. 185-200. Wasson, R. Gordon, Albert Hofmann, and Carl A.P. Ruck. Road to Eleusis. Unveiling the Secret of the Mysteries. New York and London: Harcourt Brace Jovanovich, 1978. Watelet, Claude-Henri and Pierre Charles Lévesque, eds. Dictionnaire des arts de peinture, sculpture et gravure. 5 vols. Paris: Prault, 1792. Wayman, Patrick A. “The Grubb Astrographic Telescopes.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 139-42. Wegner, Ulrich. Afrikanische Saiteninstrumente. Berlin: Museum für Völkerkunde, 1984. Weimer, Théo. “Naissance et développement de la Carte du Ciel.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 29-32. White, Graeme L. “The Carte du Ciel – The Australian Connection.” Mapping the Sky. Past Heritage and Future Directions. Proceedings of the 133rd Symposium of the International Astrophysical Union. Ed. Suzanne Débarbat et al. Dordrecht, Boston, and London: Kluwer, 1988. 45-51. Wiedeburg, Paul. Der junge Leibniz. Das Reich und Europa. Part II: Paris. Vol. I: Europäische Politik. Wiesbaden: Steiner, 1970. Wiemers, Michael. Bildform und Werkgenese. Studien zur zeichnerischen Bildvorbereitung in der italienischen Malerei zwischen 1450 und 1490. Munich and Berlin: Deutscher Kunstverlag, 1996. Wiesemann, Claudia. “Instrumentalisierte Instrumente: EEG, zerebrale Angiographie und die Etablierung des Hirntodkonzeptes.” Instrument – Experiment. Historische Studien. Ed. Christoph Meinel. Berlin: GNT-Verlag, 2000. Wiesenfeldt, Gerhard. Leerer Raum in Minervas Haus. Experimentelle Naturlehre an der Universität Leiden, 1675-1715. Amsterdam, Diepholz, and Berlin: Verlag für Geschichte der Naturwissenschaften und der Technik, 2002. Wille, Günther. Musica romana. Die Bedeutung der Musik im Leben der Römer. Amsterdam: Schippers, 1967. Williams, Robert. Art, Theory, and Culture in Sixteenth-Century Italy. From Techne to Metatechne. Cambridge: Cambridge University Press, 1997. Willis, Thomas. “De Motu Musculari.” Opera Omnia. Vol. 1. Lyon, 1681. Willis, Thomas. “The Anatomy of the Brain.” The Remaining Medical Works of that Famous and Renowned Physician Dr. Thomas Willis. Trans. Samuel Pordage. 2 vols. Montreal, 1965 [facsimile of the London edition, 1681]. Willis, Thomas. Cerebri Anatome: cui accessit nervorum descriptio et usus. London, 1664. Willis, Thomas. “Two Discourses concerning the Soul of Brutes.” Dr. Willis’s Practice of Physick. Ed. Samuel Pordage. London, 1684.
Bibliography
553
Wilson, Catherine. The Invisible World. Early Modern Philosophy and the Invention of the Microscope. Princeton: Princeton University Press, 1995. Wind, Edgar. Art and Anarchy. Illinois: Northwestern University Press, 1985. Winterfeld, Friedrich Wilhelm von. Teutsche und Ceremonial-Politica. 2 vols. Frankfurt a.M. and Leipzig, 1700 and 1702. Winterhalter, Albert G. The International Astrophotographical Congress and A Visit to Certain European Observatories and other Institutions. Report to the Superintendent. Washington: Government Printing Office, 1889. Wirth, Louis. The Ghetto. Chicago: University of Chicago Press, 1928. Wirth, Uwe, ed. Performanz. Zwischen Sprachphilosophie und Kulturwissenschaften. Frankfurt a.M.: Suhrkamp, 2002. Wise, Norton and Crosbie Smith. Energy for Empire. A Biographical Study of Lord Kelvin. Cambridge: Cambridge University Press, 1989. Witt, R.E. Isis in the Graeco-Roman World. London: Thames and Hudson, 1971. Wolf, Charles. “Recherches sur l’équation personnelle dans les observations de passage.” Annales de l’Observatoire de Paris. Mémoires 8 (1866): 188. Wolfe, Tom. The Painted Word. New York: Bantam, 1975. Wolfenbüttel Thesaurus for Early Modern Festival Books. Online source: http://www.hab.de/forschung/projekte/festkultur.htm. Wolff, Christian. “Vorrede.” Bernard Forest de Belidor. Architectura Hydraulica. Oder: Die Kunst, das Gewässer zu denen verschiedentlichen Nothwendigkeiten des menschlichen Lebens zu leiten, in die Höhe zu bringen, und vortheilhaftig anzuwenden. Augsburg, 1764. Wolfschmidt, Gudrun. “Planeten, Kometen, Finsternisse – Peter Apian, Astronom und Instrumentenbauer.” Peter Apian. Astronomie, Kosmographie und Mathematik am Beginn der Neuzeit. Ed. Karl Röttel. 2nd edition. Buxheim: Polygon-Verlag, 1997. 93-106. Wood, Paul B. “Die philosophische Ausrichtung der Royal Society.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Jean-Pierre Schobinger. Vol. 3.2. Basel: Schwabe, 1988. 391-94. Wood, Paul B. “Robert Boyle.” Die Philosophie des 17. Jahrhunderts (= Grundriss der Geschichte der Philosophie). Ed. Jean-Pierre Schobinger. Vol. 3.2. Basel: Schwabe, 1988. 395-403. Woods-Marsden, Joanna. Renaissance Self-Portraiture. The Visual Construction of Identity and the Social Status of the Artist. New Haven and London: Yale University Press, 1998. Woolf, Harry. The Transits of Venus. A Study in Eighteenth-Century Science. Princeton: Princeton University Press, 1959. Yates, Frances A. The Art of Memory. London: Routledge & Kegan Paul, 1966. Yates, Frances A. The Occult Philosophy in the Elisabethan Age. London and Boston: Routledge, 1979. Zahn, Johannes. Oculus artificialis teledioptricus. Herbipoli, 1685 and Nuremberg, 1702. Zak, Sabine. Musik als “Ehr und Zier” im mittelalterlichen Reich. Studien zur Musik im höfischen Leben, Recht und Zeremoniell. Neuss: Päffgen, 1979. Zanetti, Antonio Maria. Della pittura veneziana. Venice: Albrizzi, 1797. Zeising, Heinrich. Theatri machinarum [6 parts]. Leipzig, 1607-1614. Zorbaugh, Harvey W. The Gold Coast and the Slum. Chicago: University Press, 1929. Zuccari, Federico. Il passaggio per l’Italia con la dimora di Parma. Bologna, 1608. Zuccari, Federico. “Idea de’ pittori, scultori et architetti.” Scritti d’Arte del Cinquecento. Ed. Paola Barocchi. Milan: Ricciardi, 1973. Vol. 2, 2062-118. Zuylen, J. van. “The Microscopes of Antoni van Leeuwenhoek.” Journal of Microscopy 121.3 (1981): 309-28.
Index of Names
Ach, Johann Frantz Griendel von 342-43 Addams, Jane 406, 414 Adorno, Theodor W. 177, 496 Agricola, Georg 153, 156, 160 Airy, George Biddell 317 Albers, Josef 397 Alberti, Leon Battista 24-28, 82-84, 123, 384, 386, 474 Albertus Magnus 203 Alcmaeon of Croton 180 Aleotti, Giovanni Battista 162 Alexander VII 241 Alexander the Great 241 Algarotti, Francesco 114 Al Hazen de Basrah (Abu Ali alHasan Ibn Al-Haitham) 385-87, 390 Alsted, Johann Heinrich 43 Althusser, Louis 422 Ambrose 442 Archimedes 94, 158 Argelander, Friedrich Wilhelm August 320 Aristotle 82, 155, 181, 184, 211, 442 Aristoxenus 20 Arminius, Jacobus 226 Atwood, George 291 Aubignac, François Hédelin, Abbé d’ 63, 161, 163 Auden, Wystan Hugh 396 Austin, John L. 461 Bachelard, Gaston 3, 6, 422 Bacon, Francis 63-65, 68, 83, 204, 296-97 Bacon, Roger 88, 203, 206, 385 Baecker, Dirk 475 Baeyer, Johann Jacob 309-10
Baldinucci, Filippo 113 Baltrušaitis, Jurgis 32 Bandinelli, Baccio 86 Barbaro, Daniele 23, 31 Barthes, Roland 32, 36 Bassano, Jacopo 114 Baziotes, William 397, 399-401 Becher, Johann Joachim 206, 208-19 Benton, Thomas Hart 398 Bergson, Henri 421-31, 434-35 Berlioz, Hector 259-60 Bernard, Claude 11, 421 Bernini, Gianlorenzo 69 Bessel, Friedrich Wilhelm 320 Bettinelli, Saverio 119-20 Beuys, Joseph 180 Bloch, Marc 444 Blumenberg, Hans 449 Böckler, Georg Andreas 153, 158 Bodin, Jean 439 Boerhaave, Herman 232 Bolyai, János 35 Bolzano, Bernhard 34 Bonanni, Philippo 343, 346 Bone, Muirhead 125 Borgia, Cesare 88 Boschini, Marco 113-14, 118 Bosman, Willem 274-78 Bourdieu, Pierre 452-53 Boyle, Robert 65-66, 184, 202, 204, 206, 219, 228-29, 242 Branca, Giovanni 156 Bredekamp, Horst 88 Breidbach, Olaf 193 Britto, Giovanni 84 Bronowski, Jacob 298 Brown, Arthur 399 Brunelleschi, Filippo 22, 474-75
Index of Names
556 Buchner, Eduard 11 Bull, Lucien 430 Buontalenti, Bernardo 69, 72, 74, 162 Burdach, Karl Friedrich 56, 58 Burgess, Ernest W. 394, 407-08, 411, 414-15 Calderón de la Barca, Pedro de 70 Camillo, Giulio 44 Canetti, Elias 486-87 Caravaggio, Michelangelo 385 Carpi, Jacopo Berengario da 179, 184 Carr, William 222 Castelli, Leo 402 Castiglione, Baldassare 84 Caus, Salomon de 99, 153, 165 Cavendish, Henry 282 Cézanne, Paul 197 Chadarevian, Soraya de 12 Chagall, Marc 400 Charlemagne 438, 454-55 Charles I 447 Chevalley, Catherine 244 Christian Ernst, Margrave of Brandenburg 240 Christina, Queen of Sweden 238, 241-42 Churchill, Winston 438 Cipolla, Carlo 442, 445 Clark, Andy 416 Clement VII (Giulio de’ Medici) 85 Clerk, John 222 Coates, Robert 395 Cochin, Charles-Nicolas 120 Coiter, Volcher 184 Colbert, Jean-Baptiste 171 Colombo, Realdo 179 Common, Andrew Ainslie 305-06 Comte, Auguste 422 Copernicus, Nicolaus 94, 303 Corbin, Alain 483 Corneille, Pierre 70 Corneille, Thomas 71 Coulomb, Charles-Augustin de 283 Crabtree, William 315 Crease, Robert P. 392 Crevalcore, Pietro Maria da 118, 26768 Cristofori, Bartolomeo 267 Cromwell, Oliver 440, 448
Croster, Vittorio 335 Daguerre, Louis-Jacques-Mandé 391 Dalí, Salvador 397 Dandré Bardon, Michel-François 11011, 123 Dante Alighieri 458 Deleuze, Gilles 197, 428 Democritus 181 Denicke, C.L. 338 Derrida, Jacques 434 Descartes, René 26, 28, 31, 63-64, 70, 154, 164-68, 178-80, 182, 184, 189, 195, 204-05, 225-27, 261-62, 296, 362, 388, 439, 467-68 Dewey, John 406 Diderot, Denis 134-35, 146, 161, 194, 286 Didi-Huberman, Georges 82 Dierig, Sven 7 Donneau de Visé, Jean 163 Dornau, Caspar 86 Draper, John William 390 Dryden, John 63 Dürer, Albrecht 23-25, 30-31, 84, 385 Duits, Charles 399 Eckermann, Johann Peter 488 Edison, Thomas 430 Edward I 443 Edward III 444 Egan, Charles 402 Einstein, Albert 384, 388 Eisler, Hanns 496 Elias, Norbert 452 Ernst, Max 397 Erxleben, Johann Christian Polykarp 288 Eschinardi, Francesco 246 Euclid 20, 26, 296, 385, 390, 476 Evelyn, John 242 Fabricius, Girolamo 178 Fahrenheit, Daniel Gabriel 232 Fallopio, Gabriele 184 Faye, Hervé 311, 316 Félibien, André 75, 112 Ferdinand, Grand Duke of Tuscany 72 Ferdinand II 263 Ferdinand III 215
Index of Names Fink, Hans-Werner 355 Florencia, Francisco de 235 Fludd, Robert 44-49, 54-58, 261 Foerster, Wilhelm 317 Fontana, Giovanni da 246 Fontenelle, Bernard Le Bovier de 64, 72, 167-68, 196-97 Francesca, Piero della 27 Francini, Tomaso 162 Francis of Assisi 30 Francisco de Borja 235 Fraunhofer, Joseph von 392 Frederick II, King of Prussia 487 Freud, Sigmund 427 Friedrich, Count Palatine of Mainz 215 Furetière, Antoine 160 Furttenbach, Joseph 69 Galilei, Galileo 65-66, 239, 258, 335-36, 340, 385, 390-91 Galileo, Vincenzo 258 Galvani, Luigi 282 Gassendi, Pierre 83, 188 Gauß, Carl Friedrich 35 Gelder, Arent de 110-11 George I, King of England 101, 354 Gerber, Christoph 351 Gersdorff, Hans von 178 Gilpin, William 68, 109 Giordano, Luca 114 Giorgio, Francesco di 101 Giotto (di Bondone) 458-59 Giustiniani, Vincenzo 107-09 Glauber, Johann Rudolf 204 Goethe, Johann Wolfgang von 137-38, 147, 488 Gomarus, Franciscus 226 Gonnessiat, François 322 Goody, Jack 368, 371, 374 Gorky, Arshile 394, 399 Gottlieb, Adolph 394, 400-01 Greenberg, Clement 403 Gresham, Thomas 441 Guattari, Félix 197 Guerrero, Miguel 235 Guggenheim, Peggy 397, 414 Guitti, Francesco 69 Guston, Philip 394
557 Hába, Alois 262 Hachette, Jean Nicolas Pierre 102 Haller, Albrecht von 191 Halley, Edmond 315 Hanslick, Eduard 278 Haraway, Donna 388 Hare, David 400 Hartmann, Julius 320 Hartwig, Thomas 356 Harvey, William 184-85 Hauksbee, Francis 217 Hauser, Arnold 24 Hayek, Friedrich von 451 Hayter, Stanley William 402 Heidanus, Abraham 226-27 Heidegger, Martin 383, 389 Heintz, Bettina 348, 360 Helmont, Johann Baptista van 204 Hennig, Jochen 340 Henry, Paul 306 Henry, Prosper 306 Henry II 334 Henry VIII, King of England 444 Hermes Trismegistos 203, 241 Herschel, John 313 Hertel, Christian Gottlieb 337-38 Hilbert, David 364, 373, 381 Hilgard, Julius E. 313 Hirsch, Andreas 264 Hirsch, Adolphe 311-12, 321 Hobbes, Thomas 74, 185, 202-03, 206, 219, 261, 445-46, 450, 455 Hockney, David 383-84 Hofmann, Hans 401, 403 Holbein, Hans (the Younger) 30 Hooke, Robert 62, 64-67, 341-46 Horace 172 Horkheimer, Max 176, 195-96 Horrocks, Jeremiah 315 Huarte de San Juan, Juan 196 Huber, Jörg 348, 360 Hugo, Victor 379 Husserl, Edmund 384, 427, 450 Huygens, Christiaan 169 Huygens, Constatijn 246 Iamblichus of Chalcis 21, 129 Ignatius of Antioch 238-39 Ignatius of Loyola 235, 237-38, 251
558 Jacobi, Carl Gustav Jacob 290-91 Jacobi, Hermann Moritz 280, 283-87, 290-93 Janis, Sidney 402 Janssen, Jules 316 Joblot, Louis 344, 346 Johns, Jasper 403 Johnson, Ben 70 Johnson, Samuel 135 Jonnsen, Claus 392 Kant, Immanuel 22, 33, 176, 178, 19192, 194, 322, 362, 369-71, 476 Kapiolani, Queen of Hawaii 318 Kempelen, Wolfgang von 455 Kepler, Johannes 22, 164, 258, 262 Ketel, Cornelis 110 Khrushchev, Nikita Sergeyevich 415 Kircher, Athanasius 40, 43-44, 204, 208-09, 235, 237-44, 246-53, 261, 263-65, 267, 489, 492, 496 Klein, Felix 34-35 Kline, Franz 394, 398, 400 Kohler, Robert 9 Kooning, Willem de 394-95, 399, 417 Kootz, Sam 402 Krohn, Wolfgang 83 Ktesibios 257 Kunigunde 334 Labat, Jean-Baptiste 273-74, 277 La Bruyère, Jean de 72, 161 Lacan, Jacques 32 Lambert, Johann Heinrich 28 La Mettrie, Julien Offray de 42 Lana Terzi, Francesco 243 Lavrentev, Mikhail Alekseevich 415 Le Boë Sylvius, Franciscus de 225, 229 Lebrun, Pierre 112 Lee, Francis 401 Leeuwenhoek, Antonie van 67, 339, 341, 345 Léger, Fernand 397, 400 Le Goff, Jacques 444 Leibniz, Gottfried Wilhelm 26, 41, 43, 101, 152, 154, 163, 168-73, 206, 296, 362-66, 381-82 Lemoine, Annick 108 Leo X (Giovanni de’ Medici) 78-79
Index of Names Leonardo da Vinci 22-25, 27-29, 31, 78-79, 81-82, 84, 87-88, 90-96, 98101, 182 Leopold I 487 Leupold, Jacob 31 Leutmann, Johann Georg 338 Lévesque, Pierre-Charles 110 Lichtenberg, Georg Christoph 288-89 Liebig, Justus von 291 Liszt, Franz 292 Lobaschefskij, Nicolaj Ivanovich 35 Locke, John 206, 388, 447 Lorrain, Claude 68 Lotti, Cosimo 70 Louis III 30 Louis IX 441 Louis XIV 75, 449, 453 Ludwig, Carl 7 Luhmann, Niklas 439, 446 Lumière, Auguste Marie Louis Nicolas 429-30 Lumière, Louis Jean 429-30 Luzzi, Mondino dei 179 Machiavelli, Niccolò 93 Maets, Carel de 228-31 Maffei, Scipione 267 Maignan, Emmanuel 30 Malaspina, Taddea 85 Manetti, Antonio 22 March, Estevan 118 Marey, Étienne-Jules 8, 430, 434-35 Marquis de Sourdéac (Alexandre de Rieux) 163 Marshall, Alfred 413 Marx, Karl 440, 451 Masaccio (Tommaso Cassai) 447 Massa, Niccolò 179 Matthias I 493 Maximilian I 487 Medici, Alessandro de’ 85 Medici, Ferdinando de’ 267 Méliès, Georges 429, 434 Mendeleev, Dimitri Ivanovitch 372 Menestrier, Claude-François 160-61, 243 Menghini Romano, Niccolò 71 Menzel, Adolph von 125 Merolla, Girolamo 273, 277 Mersenne, Marin 31, 261-63, 267
Index of Names Mesnadière, Jules de la 164 Meyen, Franz Ferdinand 4 Michelangelo (Buonarroti) 31 Mill, John Stuart 298 Mondrian, Piet 397, 400 Monet, Claude 425 Monge, Gaspard 102 Montaigne, Michel de 165 Moore, Henry 204 Morgan, Thomas Hunt 10 Mort, Jacob le 229, 231 Motherwell, Robert 394-97, 399-401, 415 Mouchez, Ernest B. 303, 308 Musschenbroek, Samuel van 228, 232-33 Nadal, Jeronimo 251 Napier, John 469 Nemesios of Emesa 181 Neumann, Franz 291 Newman, Barnett 394 Newton, Isaac 210, 227, 267, 306, 309, 391-92, 447 Nicéron, Jean-François 27 Niépce, Joseph 390 Nietzsche, Friedrich 201 Nollet, Jean-Antoine 288 Ogburn, William F. 394, 408-10 O’Keefe, John 59 Oken, Lorenz 57-58 Oldenbarnevelt, Johan van 226 Oresme, Nicole 446, 449 Orlandi, Pellegrino Antonio 113, 118 Paciolo, Luca 22 Palladio, Andrea 23, 27, 162 Palmer, Vivien 407, 409, 414 Panofsky, Erwin 23 Paracelsus 204-05 Paré, Ambroise 184, 186 Park, Robert E. 394, 406-11, 414-15 Parmigianino (Girolamo Francesco Maria Mazzola) 112 Parsons, Betty 402 Paulhan, Jean 427 Peirce, Benjamin 323-25 Peirce, Charles Sanders 299, 309, 31113, 323-27
559 Penfield, Wilder 59 Perrault, Charles 121 Perrault, Claude 190 Peruzzi, Baldassare 162 Philip IV (the Fair) of France 444 Picasso, Pablo 397 Pickering, Edward 308, 323-24, 326 Piles, Roger de 121 Pindar 257 Pino, Paolo 123 Plantamour, Émile 321 Plato 127, 129-30, 141, 144, 181, 422, 434, 459 Pleticha, Heinrich 273 Pliny the Elder 84, 130 Poisson, Nicolas Joseph 164 Pollarolo, Carlo Francesco 278 Pollock, Jackson 394-95, 399-401, 415 Pontormo, Jacopo 85 Popp, Anny 98 Praetorius, Michael 258, 260, 262-63, 267, 270-72, 277 Proclus 20, 129, 144 Pure, Michel de 161-62 Pythagoras 20-21, 73, 130 Ramelli, Agostino 90, 156, 158 Raphael (Raffaello Sanzio) 112, 424 Rasmussen, Nicolas 13 Rauschenberg, Robert 403 Redi, Francesco 184 Régnier, Nicolas 107-08, 115 Reimann, Eugen 319 Reisch, Gregor 181 Rembrandt van Rijn 110, 424 Reni, Guido 112 Reynolds, Joshua 120-21, 123, 125 Riccioli, Giambattista 243 Ridolfi, Carlo 111 Riemann, Bernhard 35 Rivera, Diego 398 Robert-Houdin, Jean-Eugène 429, 434 Roberts, Isaac 305-06 Rockefeller, John D., Sr. 407, 409, 413 Rodin, Auguste 425 Rohlfs, Gerhard 278 Rohrer, Heinrich 349, 352-52, 354, 357
560 Romano, Giulio 112 Rosenberg, Harold 399, 402 Rossellini, Roberto 428 Rothko, Mark 394-95, 397, 401, 403, 417 Rotman, Brian 470, 475, 477 Rudolf II 215, 247 Sandburg, Carl 405 Sangallo, Antonio da 162 Santos, João dos 269-70, 273, 277 Scannelli, Francesco 112 Scaramuccia, Luigi 112-13 Schapiro, Meyer 396 Scheiner, Christoph 247 Schelling, Friedrich Wilhelm Joseph 57 Scheuchzer, Johann Jacob 345 Schickore, Jutta 7 Schlegel, Wilhelm Friedrich von 437 Schleiden, Matthias Jacob 4 Schön, Erhard 30 Schott, Caspar 29, 243, 246 Scrot, William 30 Searle, John 454-55 Senguerd, Wolferd 225, 231-32 Serlio, Sebastiano 162 Seyen, Arnold 230 Sforza, Lodovico 88 ’s Gravesande, Willem Jacob 231 Shakespeare, William 44, 65 Shapin, Steven 202 Shaw, Peter 217 Shils, Edward 409 Singer, Wolf 193 Siqueiros, David Alfaro 398 Small, Albion W. 394, 408 Smith, Tony 427 Snow, John 409 Socrates 181 Soemmering, Samuel Thomas 191 Southwell, Robert 241-42 Spelman Rockefeller, Laura 407, 414 Spinoza, Baruch de 206 Stahl, Georg Ernst 214-15, 217-18 Steele, Theodore C. 414 Stensen, Niels 183-84, 190 Still, Clyfford 394, 396 Stoll, Erich 355-57 Strasser, Bruno 13
Index of Names Struve, Friedrich Georg Wilhelm 320 Svedberg, Theodor 11 Swammerdam, Jan 345 Tanguy, Yves 397, 399 Tesauro, Emanuele 66, 243 Thiele, Thorvald Nicolai 307 Thomas Aquinas 203, 455 Thomas, William I. 394, 408 Tintoretto (Jacopo Robusti) 111, 114, 118, 120 Tomlin, Bradley Walker 399 Torelli, Giacomo 70, 162 Tupman, G.L. 318-19 Turing, Alan 373-74 Turner, Herbert Hall 306-07 Turner, William 427 Ulrich, Duke of BraunschweigLüneburg 487 Valentiner, Karl Wilhelm 319 Vasari, Giorgio 82, 87, 111, 113, 119, 458 Vaucanson, Jacques de 42 Vega Carpio, Lope Félix de 70 Veranzio, Fausto 156 Vermeer, Jan 25 Vesalius, Andreas 183 Viator, Jean Pélerin 475 Vico, Enea 86 Vico, Giambattista 129 Viète, François 469 Vieussens, Raymond de 183 Vigarani, Gaspare 162, 164 Villalpando, Juan Battista 240 Villanova, Arnaldus de 203 Villarceau, Yvon 311-13 Virdung, Sebastian 260 Vitruvius 70, 99, 155-56, 160, 257 Volder, Burchard de 225, 227-31 Volta, Alessandro 281-82 Vouet, Simon 115-16, 121-22 Walbeck, Henrik Johan 320 Walgenstein, Thomas Rasmussen 246 Warhol, Andy 403 Watelet, Claude-Henri 110-11, 122 Weber, Max 405, 413 William III of Orange 226, 449
Index of Names Willis, Thomas 182-83, 185-90, 194-96 Wilson, Catherine 384 Wind, Edgar 125 Wiora, Walter 262 Wirth, Louis 394, 408, 411 Witelo, Erazmus Ciolek 385 Witt, Johan de 226 Wittgenstein, Ludwig 193 Wittichius, Christoph 227 Wolf, Charles 320-21 Wolfe, Tom 401
561 Wolff, Christian 286 Wren, Christopher 182 Yates, Frances A. 44 Young, Thomas 392 Zahn, Johannes 338 Zanetti, Antonio Maria 114 Zeising, Heinrich 156-58 Zonca, Vittorio 156 Zuccari, Frederico 63, 83 Zur Lippe, Rudolf 23
Index of Subjects
Academy 152, 154, 170-72, 222-23, 233, 238, 261, 303 Academy of Sciences/ Académie des Sciences 167, 303, 316 Akademie der Repräsentationen 154, 170-72 Akademie der Wissenschaften 303 Acoustics 245, 258, 266-67, 480-87, 492-95, 497-500, 502-03 Actor 72-73, 134, 190, 247, 452 Administration 74, 194, 302, 365, 398 Admiration 113, 121, 152-54, 161, 166, 255, 269 Aesthetics 24-25, 65, 127-28, 131, 453 Affect 161, 166, 183, 495 Africa/African 269-75, 277-79, 309 Air Pump see Instrument Alchemy 47, 83, 86-87, 201, 203-19 Alien (the) 129, 133 Allegory 66, 74, 144, 166, 180, 235 America/American 16, 63, 235, 26970, 309, 314, 316, 319, 325, 327, 394, 396-97, 400-01, 404-05, 410, 414 Anamorphosis 21, 28-36, 253 Anatomical Theater see also Theatrum Anatomicum 224, 230-31 Anatomy/anatomical 5, 46, 57, 78, 90, 170, 177-83, 187, 189-90, 192, 196, 264, 282 Anatomical Cartography 187 Animal 5, 7, 57, 124, 134, 142, 160, 164, 180-82, 185-89, 209-10, 252, 282, 343, 345 Anthropology 56, 146, 164, 284, 368, 409 Antiquity 25, 84, 88, 139, 180, 203, 257-58, 372, 378, 385, 390, 499
Apparatus 2, 5-9, 17, 23, 29, 31, 6869, 71, 73, 88, 91, 128, 132, 138-39, 143, 148, 158, 160, 162-63, 166, 169, 195-96, 218, 289, 312, 314, 316, 319-21, 325, 341, 355, 434, 451, 455, 501 - optical 6, 341 Archaeology 409 Architecture 3, 26, 53, 58, 69-70, 83, 102, 154-55, 162, 168, 182-83, 187, 189, 222-23, 240, 244, 257, 391, 394, 401, 415-17, 453, 458, 474, 492, 494 Archive 6, 409 Arithmetic 20, 41, 81, 262, 296, 36263, 457, 460, 464-65, 467, 472, 476 Ars 25-26, 40, 44, 51, 53, 106, 110, 123, 172, 235, 246, 249, 252, 264, 459 Ars memoria 51, 53 Artes 81, 106 Artes liberales 25, 82, 267, 292 Artes mechanicae 25, 153 Art History see History of Art Artifact 3, 5-6, 11, 73, 146, 257, 289, 337, 352, 372, 389 Artifice 131, 173, 428, 434 Artisan/artisanal 2, 233, 267, 285-86, 290, 293 Artist 22-23, 25-26, 87, 98, 100, 10607, 109-14, 118-21, 123, 125, 134, 214, 292, 383, 394-404, 412-417, 459 Art of Experimenting 286, 290, 293 Art School 401, 403 Asia/Asian 139, 144, 309 Astonishment 64, 72, 75, 153, 160, 166-67, 255 Astrology 70, 130
564 Astronomy 167, 303, 321 Audience 42, 71-72, 75, 87, 95, 133, 138, 141, 152-54, 157, 159-60, 163, 179, 246-47, 250, 263, 386, 395, 401, 414, 488 Aura 87, 159 Authenticity 305, 402 Authority 204, 298, 308, 312, 327, 352, 439, 488 Automat/Automaton 167, 169-70 Avant-Garde 398 Barometer see Instrument Baroque 29, 32, 72, 130-32, 136, 153, 161-62, 168, 172, 266, 278, 335, 492-93, 496, 499 Biochemistry 11, 17 Biology see Life Sciences Black box 127, 136-38 Body 3, 6-7, 45-46, 48, 50, 53-54, 56, 82, 90, 99, 124-25, 127-28, 147, 164-65, 176-81, 183-84, 18687, 190-94, 209, 211, 218, 227, 229, 270-74, 284, 343, 388, 431, 434-35, 442, 450, 459, 475-76, 486, 501 Body image/schema/model 48, 50, 53, 56, 179, 183, 186-87, 191-92, 194 Border 28, 128, 132, 502 Botany 4, 228, 230, 232 Brain 56, 59, 129, 166, 178-94, 19697, 316, 389, 416 Brush see Instrument Bureaucracy 407 Cabalism 43, 203, 212-13 Cabinet 30, 232, 258 Camera Obscura 137, 245, 247, 38386, 389-93 Canvas 22, 106, 110, 116, 146, 178, 398-99, 417 Central Perspective 21-22, 24, 28-29, 32, 447, 449-50, 452-53, 459-60, 473-74, 494 Ceremonial 244, 479-82, 484-85, 48889, 492, 494-96, 498, 500, 502-03 Ceremony 146, 461, 479-81 Certainty 67, 227, 364-65, 375, 381, 467
Index of Subjects Chaos 67-68, 81, 131, 140, 147, 440 Chapel 137-38, 483, 486 Charlatan 130, 206 Chemistry 6, 11, 16, 19, 201, 203-06, 214, 217-18, 225, 228-29, 232, 422 Christianity 132, 134, 145, 444 Chromatic 337 Chronophotography 430, 435 Church 41, 131-32, 138, 222, 226, 238, 240-41, 247, 442, 444, 483-84, 486, 492, 495 Cipher 47, 466 Clock/Clockwork 41, 78, 99, 168-69, 239, 281, 284, 301, 438, 442, 449 Cogito 439 Cognition 179, 183, 192, 255, 426, 430 Collection 154, 170, 239, 244, 258, 269, 271, 335, 397, 465, 479 College 64, 222-23, 225, 250, 261, 429-30 Collegium 171 Collegium Romanum 238, 263 Colonization/Colonialism/Colonial 65, 269, 277, 301-02, 322 Color 81, 84, 86, 95, 108, 111, 11314, 116, 138, 145, 209, 211, 247, 255, 278-79, 323, 341, 386, 391, 495, 498 Combinatorics 55, 363-64, 366, 373, 375, 381 Comedy 70, 147, 170 Compass see Instrument Computation/Computational 365-66, 368, 375-76, 379 Computer 27, 193, 356, 364-66, 37375, 390, 413, 457-58, 464, 472 Consciousness 7, 128, 146, 369, 423 Continuum 463 Continuity 219, 277, 426, 428-29, 434 Control 24, 74, 113, 124, 141, 186-88, 191, 194, 216, 305, 373, 375, 459, 489, 492, 503 Corpse 182, 239 Cosmos/Cosmology/Cosmography 20, 22-23, 38, 41, 44-48, 50, 53, 5558, 132, 134, 143-44, 148, 164, 209, 264, 335, 372, 390, 400, 417, 481 Costume 142, 167, 173 Counter-Reformation 71
Index of Subjects Court/Courtly 74-75, 84, 157, 206, 258, 263, 267, 290, 297-98, 452-53, 479-82, 484, 486-89, 492, 497, 503 Craft 3, 6, 67, 69, 81, 109, 123-24, 258, 285, 290, 334 Craftsman/Craftsmanship 30, 66, 70, 74, 124, 143, 156, 160, 168, 293, 337, 355 Creation 16, 21, 26, 38, 41, 45, 48, 50, 55, 57, 64, 68, 78, 101, 111, 125, 158, 178, 187, 203-04, 209, 211-13, 239, 251, 255, 262-63, 267, 285, 288, 348-49, 360, 400, 425, 450, 461, 465, 471, 476-77, 479, 495 Cultural technique 458, 460, 465, 470-71 Cunning 129, 284 Curiositas 142 Curiosity 29, 31, 66, 142, 145, 153, 157, 166, 202-03 Dance 24, 139, 143, 407 Definition 41, 113, 155-56, 160, 176, 182-83, 195, 205, 211, 217-18, 296, 322, 476, 501 Democracy 298 Demonstration 20, 35, 55, 170, 178-79, 183-84, 208, 224, 239, 241, 243-44, 250-52, 255, 291, 398, 485, 498-99 Devil see also Satan 246-47, 322 Diagram 87, 102, 411 Dialectic 2, 6, 21, 26, 31 Digital 148, 356, 413 Dilemma 12, 148, 229, 440, 443 Discourse 7, 112, 120, 123, 125, 19697, 205, 218, 286, 369, 371-72, 379, 427, 445, 447, 450, 471 Discovery 20, 28, 33, 35-36, 63, 6567, 79, 137, 181, 183, 192, 284, 289, 307, 336, 372, 395, 459, 461, 467, 475 Disegno 83, 111, 113-14 Dissection 181, 183, 430 Distance 22, 27, 85, 90, 101, 132, 138, 141, 177, 289, 303, 309, 311, 31517, 328, 337, 350, 495, 499, 502 Distorsion 28, 32 Divine 25-26, 54, 63-64, 73, 119, 134
565 Drama 4, 24, 65, 67-68, 70-72, 137, 143, 159 Drawing Pen see Instrument Dream 133-34, 136, 143, 153, 196, 296 Dynamics 7, 11, 101 Early Modern Age 20-21, 23, 26, 34, 81, 109, 118, 128, 178, 187, 205, 208, 244, 286, 297, 385, 389-90, 392, 457, 459, 473, 476, 479-80, 482-83, 487-89, 493, 495, 498-500, 503 Economy 285, 297, 363, 451 Effect - aesthetic 68 - of perception 23, 136 - optical/visual 95, 111, 117, 133, 235, 249, 252, 341, 385-86, 390 - theatrical 31, 70, 72-73, 75, 161, 163 Electron Microscope see Instrument Emblem/emblematic 1, 5, 17, 66, 14243, 145, 160, 232, 243-44, 494 Embodiment 3, 68, 70, 75, 129, 459 Emergence 4, 193, 202, 210, 212, 303, 365, 368-69, 477 Emotion/emotional 79, 120, 134, 189, 192, 238, 367, 396 Empirical Knowledge 209, 213-14 Empiricism 26 Encyclopaedia 181, 195, 286 Engineer/Engineering 18, 26, 69-70, 74, 81, 99, 101-02, 154, 156-58, 162, 257, 268, 283-84, 286, 289-90, 293, 380 Engraving 31, 99, 237-38, 243, 24546, 249, 253, 343, 489 Enlightenment 128, 130-33, 139-40, 143, 145, 148, 176, 281, 297, 477 Epiphany 136, 140, 142, 450, Epistemology 7, 286, 299-300, 325, 388-90, 422 Esoteric 128, 145, 215, 219 Ethical 95, 148, 302 Euclidean 25, 29, 33-36 Europe/European 57, 63, 90, 128, 133, 159, 206-07, 223-24, 233, 269-70, 272-74, 277-78, 302, 308-09, 312, 324-25, 385, 387, 397, 401, 405, 441-42, 450, 454, 467, 484, 489, 496
566 Evidence 1, 84, 91, 111, 122, 141, 161, 203, 208, 215, 246, 252, 269, 273, 299, 326, 362-63, 416, 459 Exchange 299, 402, 439, 446, 448, 470-72 Exoteric 215, 218-19 Exhibition 133, 146, 170, 178, 346, 398, 415 Experience 38, 42, 44, 53, 55, 59, 62, 68, 116, 120, 124, 138-40, 148, 164, 169-70, 179, 208, 218-19, 230, 23738, 260, 269, 285, 288, 291, 293, 302, 315-17, 319, 326, 352, 369, 371, 380, 396, 415, 421-23, 427-28, 433, 435, 447-48, 476, 483 Experientia 205, 208 Experiment 3, 8-9, 11-12, 22, 28, 59, 67, 69, 83-84, 170, 179, 183, 18586, 192, 195, 201-05, 207-10, 21216, 218-19, 224-25, 227-30, 232-33, 239, 241-42, 245-47, 252, 258, 267, 281, 283, 286-89, 291, 296, 305, 320, 322, 349, 351-52, 356-57, 35960, 388, 391-92, 413, 431, 469, 474 Experimental Art 26, 66 Experimental Knowledge 287, 28991, 293 Experimental Method 32, 201-02, 204, 207-08, 213, 217, 421-22 Experimental Observation 66 Experimental Practice 201, 203, 209, 219 Experimental Set-Up 190 Experimental Science 1, 202-03, 229-30, 233, 287, 292 Experimental System 2-3, 6, 9, 17, 290 Experimental Techniques 201 Experimentalization 7 Experimentation 3, 195-96, 201, 205, 208, 215, 218, 277, 282, 308, 348, 380, 400 Expert Knowledge 124-25 Expertise 128, 148 Exposition 302 Eye 4, 11, 22-23, 34, 66-68, 82, 95, 100-01, 130, 132, 135, 141, 158, 164, 167, 171, 180, 183, 213, 24647, 297, 310, 316, 341, 364, 369, 386, 388, 435, 474-75
Index of Subjects Faith 62, 93, 243, 250, 364 Fantasy 134, 188, 212, 279 Fascination 36, 65, 67, 71, 75, 203 Festival 139, 480, 485, 492 Fiction/Fictionalism 74, 135, 141, 161, 172, 243, 245, 454-55 Figuration 53, 95, 167, 301 Film 401, 428-29, 434, 496 Firework see Pyrotechnics Flying Machine see Machine Fortification/Fortress 69, 74, 88, 90, 121, 158, 187, 493, 498-99 Framework 2, 9, 25, 96, 224, 328, 494 Fresco 28, 138 Function 1, 4-5, 9-12, 16, 36, 42, 5051, 53, 62, 69, 90, 99, 109, 131, 152-53, 164-65, 172, 177-81, 18391, 193, 195-96, 208, 219, 224, 230, 232, 243, 334, 375, 377-78, 386, 411, 463, 465, 467, 471-72, 476, 479, 492, 494-96, 500 Future 4, 137, 169, 227, 241, 302, 304, 372, 374-75, 404, 440-41, 443 Gallery 170-71, 203, 397, 401-03, 415, 417 Garden 75, 164, 180, 223, 230, 402, 489, 493, 495 Gravitation 11, 302, 309, 328 Generation 16, 121, 141, 162, 181, 206, 210-13, 287, 305, 397, 401 Genius 26, 87-88, 110-11, 114, 11922, 289 Geodesy 303, 326 Geography/geographic 94, 132, 144, 274, 278, 300, 302, 394-95, 398, 401, 407, 411-15 Geology 83 Geometry/geometric 20-36, 70, 81, 86, 88, 90, 99, 132, 164, 267, 289, 362-63, 390, 431, 435, 459, 467, 469, 472, 477 Geometricalization 23, 25 Gesture 81, 95,128-29, 157-58, 240, 431 Glasses see Instrument Globalization 130 Globe 44, 144, 245, 253, 301, 413, 438 Globe Theater 44
Index of Subjects God/Goddess 26, 45-49, 54-56, 58, 63, 73, 95, 127-30, 133, 135-36, 140, 142-44, 146, 148, 162, 212, 237-39, 243-45, 262, 345, 388, 450, 455, 489, 499 Government 88, 95, 194, 227, 318, 327, 407, 413, 452 Grotesque 132, 144 Gunpowder 94 Habit/habitual 121, 301, 326, 401 Hand 2, 4, 81, 84, 100, 106-14, 11718, 121-24, 128, 133, 135, 142, 15758, 176, 238, 270, 288, 297, 305-06, 322, 383, 395 Handicraft 2 Harmony 25, 53-54, 69, 73, 262, 266, 269, 481 Hearing see Senses Heart 31, 49-46, 86, 112, 148, 178, 180, 184-85, 189, 193-94, 235, 23740, 246, 249, 255, 489 Heresy 239, 404 Hermeneutics 96, 128 Hermetics 131, 138, 203-04, 207-08, 216, 241, 417 Hierarchy 43, 48, 57, 112, 134, 181, 183, 194-95, 205, 249, 297, 372, 376, 479 Hieroglyphs 142, 147, 243 History 4, 16, 23, 30, 35, 42-43, 70, 75, 80, 87, 101, 113, 127-28, 130, 134, 141, 146, 148, 202, 210, 219, 224, 244-45, 257, 260, 277, 280, 288, 366-68, 390, 392, 395, 438-39, 441, 449, 451, 469, 475, 501 History of Art 35, 106-07, 113, 115, 124, 128, 417 History of Knowledge 286, History of Science 201-02, 218-19, 348, 461 Humanism 73, 299 Hybrid/Hybridization 9, 135, 144, 148, 293, 458, 461, 463, 471-72 Hydraulics 69, 87, 160, 162, 180, 186, 191, 239, 245 Iconography 145, 244-46, 250 Idea of man/Image of man 55, 63
567 Ideal/Idealization 25-26, 70, 205, 208, 230, 233, 273, 293, 305, 348, 494 Illusion/Illusionism 21-22, 98, 116-17, 128-29, 133-38, 143, 153, 161-62, 169-70, 172, 245, 251, 428-29, 431, 434-35, 448, 450-54, 474-75 Image 12, 22, 26-27, 29-31, 38-40, 4358, 70, 75, 80, 82, 84-85, 90, 101, 122, 128, 132, 135, 141-43, 146, 206, 223, 241, 243-47, 249, 252, 302, 348-60, 383-86, 388, 390-92, 400, 422-31, 435, 439, 445, 452, 454, 459-61, 463-64, 474, 494 Imagination 53, 56, 112, 116, 120, 135, 139, 152, 162-63, 181, 188-89, 237, 244, 249-50, 341, 369-71 Imitation 85, 98, 111, 121, 124, 202, 212, 245, 263, 281, 499 Immersion 396 Impostor 404 Improvement 64, 67, 196, 241, 503 Incommensurability 467 India/Indian 142, 147, 313, 465-68 Individual/Individualism 74, 132, 152, 170, 202, 240, 301, 303, 307-08, 324-27, 378, 415, 500 Industry/Industrialization 2, 90, 121, 145, 196, 285-86, 290, 298, 302, 354, 401, 405-06, 410, 442 Information 12, 46, 55, 193, 241-42, 299, 341, 348, 407, 413, 501 Ingenium 87, 156, 158 Installation 489, 492, 494 Instrument 1-4, 6-7, 9-11, 14-19, 2425, 62, 64-68, 75, 78, 80-81, 101, 106-09, 112, 114-15, 122, 124-25, 132, 152, 156, 176-80, 182, 184, 186, 194, 196, 216, 224, 228, 23033, 257-68, 270-77, 289, 292, 294, 300-01, 305-06, 308, 310-12, 31416, 318, 325, 327, 335, 337-38, 34042, 348-49, 351-52, 360, 366, 387, 391-92, 394, 412, 423, 425, 428, 430, 434-35, 437-38, 442, 448-49, 453, 457-59, 463-65, 479-82, 485, 487-88, 492, 494-99, 501-03 - calculating 457, 464-65 - optical 66, 75, 148 Air Pump 228, 230, 232 Barometer 230, 284, 297
568 Brush 84, 95, 106-124, 142, 383-84, 458 Compass 25, 458-59 Drawing Pen 81, 84 Electron Microscope 2, 13-14, 1617, 349 Glasses 66-68, 79, 335 Lens 4, 6, 28, 68, 131, 133, 180, 246, 252, 334-47, 349, 386 Microscope 4-7, 12-14, 16, 62, 65, 67-68, 232, 252, 337-38, 340-43, 346, 348-52, 354-58, 360, 384 - musical 124, 257-64, 266-69, 271-72, 275-76, 480-82, 496-97, 501, 503 Pendulum 311, 309, 325 Phenakistiscope 252 Reflecting Telescope 305-06, 325, 391 Ruler 25, 78 Telescope 62, 66-68, 323, 235, 305-06, 318, 335, 387, 390-91, 449, 458 Thermometer 297, 442 Ultracentrifuge 11 Instrument Maker 2, 8, 218, 228, 232, 258-59, 267, 289-90, 312 Instrumentalism 462 Instrumentalization 23, 75, 127, 132, 176-78, 194-96, 367-68, 371, 429, 464, 502 Instrumentation 29, 260, 303, 307, 314, 323, 384, 389 Instrumentology 479, 482 Interaction 6, 14, 40, 165, 169, 178, 208-09, 280, 283, 293, 349, 352, 360, 376-77, 461-62, 503 Internationale Gradmessung 299, 302-03, 315, 327 Intuition/Intuitionism 177, 251, 36264, 369, 371, 424, 426-27 Interface 133, 143, 148 Inventory 219, 378 Investigation 2-3, 7, 10-12, 14, 21-22, 26, 107, 112, 178, 229-30, 257, 28790, 294, 297, 307, 351, 378, 381, 411, 471, 501 Invisibility/(the) Invisible 7, 95, 133, 212, 460
Index of Subjects Jesuits 42, 119, 128, 131, 160, 237-38, 242-45, 250-52, 261, 263 Judgment 189, 191, 195, 269, 279, 301, 307, 319, 324 Juggler 133 Know-how 95 Knowledge 3, 5, 30, 38, 43, 55, 64, 67, 69, 81-82, 114, 124-25, 148, 179-80, 181, 188, 205-06, 208-09, 213-14, 224, 230, 241, 250, 260, 262, 264, 267, 269, 285-91, 293, 323, 345, 351, 360, 371, 378, 380, 384, 386-89, 406, 411, 422-24, 435, 442, 461 Knowledge organization 43, 58 Knowledge production 287, 388-89 Knowledge representation 38 Knowledge tradition 290, 293 Kunstkammer see Wunderkammer Laboratory 7, 10, 16-17, 63, 80-81, 87, 176, 195, 197, 201, 206-07, 21517, 219, 223-25, 228-32, 302, 349, 354, 356, 358, 360, 410-13, 415, 417, 421, 461 Labyrinth 140, 239, 376 Landscape 29-30, 33, 40, 48, 59, 68, 95-96, 99, 134, 197, 247, 300, 39192 Laterna Magica/Magic Lantern 131, 133, 136, 146, 170, 245-46, 249, 252 Lens see Instrument Lexicography 159 Library 43, 182, 223, 267, 379, 408 Life Sciences 2-5, 7, 9, 12, 15-18, 83, 303, 421 Light 14, 17, 54-55, 67-69, 83, 86, 95, 108-09, 116-117, 124, 138-42, 144, 146, 235, 241-42, 246, 252-53, 255, 281, 310, 319, 338, 385, 387, 391-92, 399, 411 Literature 1, 4, 74, 111, 120, 153, 155, 270, 399 Logic 21, 24, 27-29, 34, 41-43, 169, 219, 230, 261, 263, 372, 413, 442-43, 445, 453, 458, 464, 470-72
Index of Subjects Machine 7, 9, 15-16, 29, 31, 41-42, 62-63, 68-75, 90, 99, 101-02, 130, 152-70, 172, 180, 186-87, 191-92, 194-96, 245, 247, 250-51, 280, 28387, 290-91, 293-94, 318, 320-22, 324, 339-40, 351, 365, 373-74, 430, 438-39, 450-52, 454-55, 457-58, 464, 494, 499, 501, 503 Flying machine 70, 87, 94, 160 Theater machine 69, 71-75, 159-63 Macrocosm 38, 55, 57, 298 Magic 25, 36, 53, 75, 86, 127-37, 139, 141-42, 144, 146, 148, 161-62, 170, 213, 239, 246-47, 249-50, 266 Magnet/Magnetic 133, 239-41, 245, 280, 282-84, 287, 289-91, 293-94, 397, 413 Mannerism 31 Manuscript 79, 87-88, 92, 99-102, 162, 215, 284, 324, 326-27 Map see also Cartography 9, 38, 48, 50, 56, 83, 94, 136, 235, 300, 30203, 305-07, 308-09, 314, 322-23, 328, 407, 409, 438 Mask 30, 247, 255 Material/Materiality 3-4, 7, 10, 14, 41, 47, 56, 65, 75, 81-82, 84, 90, 95-96, 109, 116, 122, 128, 140-42, 145, 155, 163, 169, 176-77, 179-80, 19192, 194, 196, 203, 209, 211-13, 21617, 237, 264, 266, 285, 302, 338, 360, 363, 366-68, 379-80, 395, 405, 413, 415-16, 422, 442, 457, 465, 470, 472, 492 Mathematics 12, 20-22, 24-29, 31-36, 41, 44, 162, 231, 245, 261, 263-64, 267-68, 290-92, 309, 311, 323, 36366, 372-74, 381, 415, 431, 447, 457, 459-60, 462, 464-65, 467, 469, 471, 473-76, 481 Measure 9, 11, 16-17, 24, 38, 45, 48, 53, 55-56, 58, 65, 80, 110, 130, 189, 193, 196, 216, 219, 226-28, 258, 277, 285, 287, 301, 309-10, 312, 315-17, 320, 325-28, 350, 366, 396, 410, 417, 426, 429, 431, 443, 459, 467-68, 475, 481, 488 Measurement 3, 7, 16-17, 20, 24, 79, 289, 301, 305-10, 312-17, 320, 32326, 348-49, 351, 354-56, 360
569 Mechanics 9, 22, 25, 29, 42, 47, 55, 69, 80, 87-88, 99-101, 106, 121, 123, 125, 141, 152-54, 156-57, 159, 162, 164-65, 168-70, 172, 184, 187, 191, 196, 204, 228, 232, 267-68, 280-81, 283-84, 286, 290, 293, 305, 307, 326, 349, 367, 425-26, 429-30, 434, 438, 449, 457, 472, 495 Mechanism 42, 47, 72, 74, 122, 128, 132, 148, 152, 155, 158, 168, 183, 191, 196, 250, 281, 350, 429-30, 435, 448, 452-54, 457, 477 Media 1, 8, 40, 69, 74-75, 94-96, 120, 125, 127-28, 132-33, 141, 146, 148, 192, 210, 317, 388, 402, 452, 45758, 463-66, 469, 471-74, 476-77, 479, 484, 500 Medialization 464 Mediation 8, 40, 227, 229, 235, 241, 244, 286, 290, 348, 362, 367, 386, 413, 442, 471-72, 475, 477, 493 Medicine 16, 47, 70, 83, 130, 172, 192, 204, 206, 224-25, 228-32, 261, 311 Meditation 64, 66, 167, 243 Medium 8, 17, 40, 69, 74, 81-83, 9596, 113-14, 116, 122, 125, 129, 192, 205, 210, 457-58, 463-66, 469, 47174 Memoria 51, 53, 430 Memory 45, 53, 56, 58, 64, 67, 124, 137, 147, 181, 188-89, 288, 341, 367, 373-74, 377, 379, 417, 422-23, 429-31 Metamorphosis 132, 142, 252, 444 Metaphor 9, 12, 62, 66, 118, 128, 164, 240, 323, 335, 374, 387-90, 498 Metaphysics 164, 169, 227-28, 261, 301, 422-23, 426, 434-35, 446, 450, 455 Method 21-22, 26-27, 31-32, 50, 6566, 99, 164, 181, 186-87, 189, 192, 201-02, 204-05, 207, 209, 212-19, 250, 260, 288-89, 291, 302, 305-08, 315-16, 327, 351, 354-55, 364, 371, 378, 390, 409-10, 421-23, 430, 434, 442, 445, 447, 459, 469 Methodology 5, 7, 208, 243, 409-10, 423
570 Microcosm 38, 51, 54-55, 57, 69, 74, 112-13, 146, 298, 345 Microscope see Instrument Electron Microscope see Instrument Middle Ages 25, 41, 155, 203, 378, 441, 445, 450, 455, 483 Military 69, 87, 90, 102, 171, 269, 302, 309, 315, 327, 449, 485, 488, 493-94, 496 Mimesis see also Imitation 25, 95-96, 98, 109, 116-17, 177 Minerals 145, 210, 217, 296 Miracle see also Wonder 26, 29, 32, 35, 75, 88, 127, 130, 134, 141, 155, 239, 243, 251, 266, 402, 441-42, 454 Mirror 22, 29, 31, 36, 66, 68, 108-09, 127, 131, 133, 142, 235, 245-47, 251-52, 255, 391-92, 438, 474-75, 494 Mission 244, 269, 277, 299, 317-18, 327 Mnemotechnique 430 Model 9-10, 13, 19, 27, 34-35, 38-43, 45, 56-57, 68, 72, 75, 98-99, 102, 109, 123, 160, 170, 178-81, 186, 189, 192-94, 206, 227, 230, 233, 238, 240-41, 263, 270, 274, 280-84, 290-91, 297-98, 302, 319-20, 335, 346, 350, 356, 365, 388-89, 397, 408, 410-12, 435, 439-40, 459, 47980, 484, 493-94, 503 Modernity 12, 38, 40-41, 59, 95, 132, 136, 146-47, 155, 201-04, 217-18, 260, 267-68, 277, 298, 379, 383, 397, 403, 410, 428, 435, 438-39, 443, 447, 458-59, 475-76, 498 Monad 132, 169 Monastery 30, 222, 261 Money 78, 215, 247, 307, 407, 409, 437, 439-55 Monk 273 Monochrome 143-44 Monstrosity 10, 29, 132, 166, 179 Motion 70, 72, 100, 113, 136, 147, 155, 169, 239, 280, 304, 396, 472 Movement 8, 16, 88, 90, 99-101, 113, 128-29, 136, 155, 160-61, 166-67, 169, 180, 184-85, 190, 194, 196, 252, 267, 273, 280-81, 283-84, 316,
Index of Subjects 350, 358-59, 366, 396-97, 421, 42331, 434-35, 470, 481, 485, 493, 501 - mechanical 267 Movie see Film Museum 237-44, 250-53, 255, 273, 334-35, 396-97, 401, 413, 415, 417 Music 70, 73, 124, 164, 257-64, 26770, 273-78, 292, 480-82, 484-85, 487-89, 493-94, 496, 498, 503 Musical Instrument 124, 257-64, 26669, 275-76, 480-82, 496-98, 501, 503 Mystery 30, 139, 243 Mystery cult 134, 141, 146 Mythology 70, 144, 244, 435, 480 Nature 3-5, 9, 11, 19-20, 23, 25-26, 38, 53, 56-58, 62-64, 66-68, 72, 75, 78, 81-83, 96, 122, 127, 133, 15356, 159, 161, 167-69, 172-73, 179, 195-96, 202-03, 205, 208-13, 216, 229, 232, 239, 243, 261, 263, 275, 281, 284-86, 288-89, 291, 293, 29698, 301, 305, 307, 337, 379-80, 39192, 406, 442, 446, 452, 454, 459, 481, 486, 495 Natural Law 164, 481 Natural History 4, 130, 145, 224, 288 Natural Philosophy 57, 202, 205-10, 224, 227-29, 231-32, 235, 237, 239, 243-45, 251, 280, 287, 296, 309, 385 Natural Science 41, 201, 204, 210, 223, 258, 283, 442 Neo-Platonism 129, 204 Nervous system 8-9, 164, 178, 180, 182, 185, 187, 189-92, 194, 316, 322 Notation 43-44, 263, 441, 463 Nothingness 467, 471, 476 Objectivity 25, 196, 305, 307, 322, 326, 348, 380, 406, 410, 417 Observation/Observer 4-5, 20, 23, 30, 33, 66, 101, 108-09, 141, 145, 181, 187, 203, 205, 209, 214, 218, 230, 246-47, 262, 291, 300-01, 303, 30506, 308-10, 312-24, 326, 328, 338, 341-43, 348, 360, 367, 386, 389, 408-11, 446, 460-61, 463, 473-75, 487, 492, 501-02
Index of Subjects Observatory 147, 223-24, 231, 303, 306-08, 317, 321, 324, 327 Ontology 128, 131, 219, 280, 389, 477 Opera 72, 147, 163, 167-68, 173, 258, 278-79, 487, 493 Operation 41, 90, 93, 122, 134, 153, 176-79, 181, 188, 216, 232, 289, 296, 301, 306, 317, 363, 365-66, 368-69, 377-78, 413, 461, 464, 467, 472, 477, 500 Optics 6, 22, 28, 30-32, 36, 66, 83, 127-28, 130, 134, 144, 148, 235, 237, 245, 247, 250, 322, 335, 337, 339, 341, 349, 384-87, 390-91 Optical Instrument see Instrument Organ see also Musical Instrument 257-60, 266-67, 270-71, 492-95, 497-99, 501 Orient 128, 130-31, 141, 144, 263, 269 Ornament 116, 121, 372 Orthodoxy 132, 204 Otherworldliness 139 Paganism 132, 138, 143, 145, 239, 244, 247, 251, 480 Painting 24-28, 30, 63, 68, 81-82, 8486, 95, 106-25, 130-31, 134, 13738, 145, 244, 383, 395-400, 402-03, 412, 416-17, 427, 459 Paper 8, 41, 86, 95-96, 98, 193, 247, 302, 343, 356, 366, 373-74, 386, 392, 448, 467, 476 Paradigm 26, 35, 80, 82-83, 201, 370, 427, 430 Paradise 63-64, 68, 73 Paradox 9, 16, 20-21, 32, 35, 67, 213, 368, 373, 377, 381, 425 Passion 86, 113, 119, 144, 166, 194, 240, 245, 251-52, 382 Pathos 78, 86 Pattern 29, 40-41, 43-44, 50, 55-58, 188, 194, 210, 219, 275, 277, 299, 351, 409-10, 412, 475, 477 Pendulum see Instrument Perception 1, 22-24, 29-30, 34, 40, 68, 125, 135-36, 153, 165-66, 169, 172, 176-77, 179, 208-09, 243, 246, 252,
571 255, 278, 320-21, 327, 370, 378, 422-23, 426-27, 429, 434, 459-60, 474 Performance 69, 73, 120-21, 125, 131, 137, 152-53, 157, 161-63, 170, 177, 218, 250, 252, 291, 486-88, 501 Performativity 128, 145-46, 201, 46162, 464, 472, 476 Personification 137 Perspective 21, 23-24, 26-36, 66, 83, 86, 88, 90-91, 95, 99, 101, 131, 15253, 158, 170-71, 177, 208, 210, 21819, 250, 284, 303, 349, 375, 386, 390, 443, 446, 459-60, 462, 470, 473, 475-77 Phantasmagoria 74, 87, 94, 128, 132, 135, 138, 146 Phenakistiscope see Instrument Philosophy 83-84, 128, 143-44, 167, 179, 192-94, 196, 204, 209, 211, 213, 216-18, 226, 229, 231, 237, 246, 262, 290, 362, 364-65, 379, 387, 421-23, 428, 431, 434, 449, 479 Photography see also Chronophotography 6, 280, 303-06, 308, 315-17, 320, 350, 390-91, 431, 434 Physician 86 Physico-Theology 345 Physics 12, 18, 28, 90, 95, 108-10, 116, 123-24, 128, 144, 152, 154, 156, 164-65, 171, 185, 189-90, 192, 195-96, 209, 212, 223-25, 227-32, 258, 261, 263-64, 267, 280, 282-84, 286, 288-89, 291, 297-98, 307, 309, 313, 328, 349-51, 355-56, 366, 37980, 392, 409, 414, 422, 430, 435, 439, 458, 460, 465, 480 Physiology 7, 9, 42, 59, 117, 181, 192, 197, 310, 321-22, 430-31 Planet 73, 147, 280-81, 284, 301, 30304, 309, 314, 316, 318, 320, 328 Play 22, 57, 70-72, 74, 88, 124, 146, 167, 170, 172, 180, 245-46, 250-51, 257-58, 260, 263, 267, 269-70, 27277, 279, 322, 452, 454, 485, 487, 496, 501 Pneumatic 62, 155, 160, 239, 245, 254 Poetics 32, 62, 66, 68, 88, 112, 119, 141, 164, 172, 314
Index of Subjects
572 Poetry 65-66, 74, 112, 119, 138, 172, 249, 398 Politics 30, 62, 65, 68, 75, 85, 148, 187, 194, 196, 203, 206, 277, 29798, 308, 327, 398-99, 404, 407, 409, 415, 437, 440, 445, 449-51, 453, 484-85, 489, 492, 500, 503 Polymath 205-06, 218, 237, 261 Power 6, 16, 34, 55, 62, 70, 72, 75, 116, 118, 128-29, 131, 133, 137, 141, 143, 148, 153, 156, 160-61, 163, 165, 188-90, 196, 208, 211, 235, 238, 244, 249, 327-28, 337, 362, 366, 370, 381, 391, 407, 42526, 446-47, 451-54, 461, 477, 479, 487, 489, 498-500 Precision 291, 306-08, 313-14, 318, 365, 423, 425, 501 Pre-modern 106, 123 Presence 17, 108, 128, 136, 143, 148, 166, 251, 364, 415-16, 422, 441-43, 452, 465, 473 Presentation 23, 40, 43, 71, 107, 138, 152, 170, 207, 215-16, 246, 250, 264, 335, 461, 464-65, 472, 486 Prima Materia 211, 213 Probability 243, 245, 310-11, 314, 317, 320, 323-24, 349, 415 Procession 131, 138, 140, 142-43 Production/Productivity 1-3, 16, 24, 30, 65, 67, 83, 91, 93, 101, 110, 158, 168, 195-96, 202-03, 212, 217, 228, 250, 287, 335-37, 340-41, 348, 356, 360, 381, 388-90, 392, 394-95, 398, 408, 412-17, 450, 452, 457, 463, 471-72, 477, 479-80, 498-99, 501 Profile 45, 116, 356, 497 Program 18, 45, 55, 64, 66, 90, 157, 194, 228, 291, 365, 373-77, 381, 408, 430, 458, 464 Progress 35, 132, 140, 144, 184, 210, 321, 356, 367-68, 398, 406, 408, 444, 469 Project 13, 16, 18-19, 22, 65, 88, 93, 95, 101, 128, 172, 204-06, 213, 286, 302-05, 307, 314-15, 323, 407, 442 Projection/Projective Technology 2122, 24, 26-27, 127-28, 131-33, 135-
36, 138-39, 245-47, 345, 390, 47576 Proof 45, 186-87, 209-10, 212, 223, 233, 354, 441, 466 Prophecy 55, 63, 88, 197, 240, 437, 451 Proportion/Proportionality 20-21, 2425, 38, 73, 82, 164, 169, 211, 263, 304, 340, 390, 473 Proscenium 158 Prototype 146, 240, 243, 439 Provisional 222, 424 Psychology 24, 136, 177, 206, 322, 326, 422, 439, 445, 495 Public/Publicity 42, 72, 88, 140-42, 148, 163, 170-72, 176, 178, 195, 215, 237, 283, 290-91, 349, 384, 402, 415, 486, 488-89, 494, 499 Public Sphere 171 Punishment 197, 227, 249, 345, 453 Pyrotechnics 69, 170, 250 Pythagorean 20, 481 Quarantore 71 Rarity 101, 170, 275, 339 Rationality 20-22, 25, 29, 31-32, 34, 55, 124, 130-31, 144, 148, 166, 181, 187-90, 192, 202-03, 326, 368-69, 374, 376-77, 379, 438, 445-46, 455 Reason 29, 31-32, 64-65, 67, 124, 161, 169, 176, 178, 180, 183, 192, 195-96, 202, 213, 261, 324, 326, 335, 362-65, 368-69, 371-72, 37479, 381-82, 455, 496 Reflection 29, 31-32, 64-65, 67, 124, 161, 169, 176, 178, 180, 183, 192, 195-96, 202, 213, 261, 324, 326, 335, 362-65, 368-69, 371-72, 37479, 381-82, 455, 496 Reflecting Telescope see Instrument Registration 7, 16, 170, 311-12, 322 Reliability 7, 269, 301, 315, 341-42, 410, 413, 440, 447 Relic 87, 334, 336 Religion 30, 71, 127-28, 130-31, 134, 136, 138-39, 143-44, 146, 148, 237, 244-45, 334-35, 403, 410, 453 Renaissance 24, 26, 74, 78, 87, 20304, 298, 383-87, 389-90
Index of Subjects Representation 3, 6, 14, 20-23, 26, 28-32, 34-36, 38, 41, 44-45, 54-55, 58, 69, 75, 81, 86, 91, 96, 98, 116, 122, 128, 131, 135, 152, 154, 16972, 180, 183, 224, 227, 229-33, 245, 251, 257, 259, 266, 272, 275, 278, 286, 308, 319, 337-38, 343-46, 348, 356, 369, 374-75, 377-78, 400, 42122, 424, 427, 453-55, 457, 459-60, 463, 466, 479, 489, 495, 500 Reproduction 5, 48-50, 56, 95, 114, 124, 146, 259, 369-70, 428 Res Cogitans/Res Extensa 164 Research Technology 2-3, 6-7, 9, 12, 18-19, 26, 83, 179, 193, 196, 230, 283, 286, 288-89, 291, 297, 381, 461-62 Resistance 14, 15, 292, 307, 403, 425, 467 Revelation 26, 62-63, 67, 75, 139-41, 143, 205, 292 Rhetoric 69, 87-88, 90-92, 100-02, 228, 242, 302, 356, 377-80, 382 Rhythm 275, 278, 322, 396, 424, 431, 500 Ritual 128, 130-32, 134, 137, 139, 143-46, 148, 270, 276, 278, 443 Rome 30, 87, 133, 179, 235, 238, 24042, 258, 263-64, 444 Rotation 8, 252, 280, 324, 339-40, 360 Royal Society 65, 204, 227-28, 241-42 Ruler see Instrument Sacrament 71, 133 Satan see also Devil 136, 251 Scenario 78, 133, 139 Scene 23, 27, 29, 62, 69-75, 90, 109, 118, 135-36, 138, 146-48, 245, 250, 257, 370, 399, 401, 404, 416, 475, 482, 484, 486, 489 Scenery 70, 74, 162, 400, 495, 497 Scenography 69 Scent see Sense Scepticism 204 Scholasticism 266, 442, 445 Scientific Organisation 167, 261, 29799, 302-03, 307, 316, 409, 413, 417 Sculpture 63, 66, 82, 86, 123, 137, 146, 494
573 Secret 20-21, 30, 129, 143, 160, 205, 208, 213, 216, 218, 261, 266, 298, 339, 341, 355, 384, 393 Sense/sensual 1, 34, 53, 55, 62, 64-65, 67, 82, 124, 128, 132, 134-36, 140, 161-62, 166-67, 176, 180-81, 192, 195, 202, 212, 237, 246, 249-52, 257, 316, 324, 327, 360, 367, 378, 380-81, 422, 431, 435, 460, 479-80, 501-02 Hearing 67, 181, 258, 262, 264, 274, 322, 369, 481, 502 Scent 67, 181, 249, 502 Sense of Touch 67, 502 Taste 67, 502 Vision 23, 28, 31, 34, 67, 100, 128, 131-32, 181, 237, 246, 249, 300, 337, 383, 390, 459, 502 Sex 56, 398, 447 Shadow 26, 95, 109, 116, 131, 133-34, 136-37, 140-41, 143, 145, 246-47, 343, 350, 358, 461 Shadow Theater 170-71 Sign system 373, 400, 465, 467, 477, 481 Simulation 40-41, 131, 135, 137, 141, 147, 380, 503 Skeleton 86, 494 Smell see Sense Society 62, 74, 132, 140, 170, 235, 239, 244-45, 290, 296, 298, 323, 325, 398, 405, 410, 438, 442, 44647, 452-53, 487 Solomon 238, 240 Solomon’s House 64, 297 Sound 120, 129, 257-58, 260-62, 26971, 273-75, 277, 311, 370, 411, 47980, 482-83, 485-86, 488-89, 492503 Space 6, 11, 18, 21, 23-25, 32-26, 44, 47-49, 56, 63-64, 68, 75, 81, 95, 98, 128, 131, 135-36, 168, 170, 177-78, 192, 222-23, 262, 273, 283, 285, 300, 302-03, 315, 343, 349, 367, 369-72, 374-78, 381, 394-95, 397, 399, 403, 409, 411-14, 425, 427, 429, 431, 435, 459-60, 463, 471-73, 475, 477, 479-80, 482-83, 486, 489, 492-96, 497, 499, 502-03 - visual 459-60
574 Species 186, 188, 300, 302-03, 376, 501 Specimen 2, 4-6, 13-14, 17, 145, 299, 341 Spectacle 69-70, 72, 74, 131, 134, 138, 142, 152-54, 161-63, 165-66, 168, 170, 239 Spectacular 70, 75, 145, 153, 159, 162-63, 170, 280-81, 321, 405 Spectator 31, 101, 116, 133-35, 140, 250, 252, 335, 496 Spectroscopy 307, 351, 391 Sphere/spherical 31, 56, 69, 73, 130, 183, 266, 337-39, 360, 387, 440, 479-81, 496, 502 Spirits of Life/Spiritus Animales 165, 180 Stage 42, 44, 47, 50, 53, 55-58, 62, 66-67, 69-70, 73-75, 110, 135-37, 141, 144, 147, 153, 158, 160, 164, 167, 173, 194, 239, 249-50, 278-79, 356, 375, 445, 453, 474, 484, 500 Stage Architecture 69 Stage Decoration 70-71, 160, 162, 244 Stage Machinery 72-74, 153, 160, 162 Stage Technique 69-70, 73-74, 250 Staging 62, 67, 71-72, 134, 153, 159, 483-85 Stoicism 203 Strategy 90, 101, 129, 170, 205, 246, 273, 293, 341, 360, 378-79, 450, 479, 484, 493 Subculture 299 Superstition 128, 130-31, 135, 148, 251 Surgery/Surgeon 124, 186 Symbol/Symbolism 35, 67, 129, 132, 140, 142-43, 145-46, 160, 177, 231, 233, 240, 245, 255, 362-66, 370, 373-75, 379-81, 383, 431, 442, 445, 452-53, 457-62, 464-65, 469-70, 472, 476, 496 System/Systematization 12, 17, 20-36, 38, 43, 57, 59, 66, 70, 82, 96, 98, 110, 123, 128, 132, 143-44, 155, 164, 168, 171, 176-77, 179, 184-86, 192, 196-97, 258, 260, 262, 267, 277, 291, 301, 303, 309, 315, 323,
Index of Subjects 326, 328, 354, 364-65, 372, 374-75, 377, 379-81, 385, 400, 407, 410, 413, 421-22, 425, 439-41, 444-47, 449-50, 453-55, 458, 462-67, 47173, 475-76, 479-80, 482, 489, 496, 498, 500, 503 Tableau 38, 43, 48, 50, 128, 134, 13738, 142-44, 146, 148 Taste see Sense Taxonomy 376 Techné 26, 83, 123, 459 Technology 2-4, 6, 9, 12, 14-18, 24, 27, 35-36, 41, 63, 69-70, 79, 81, 8788, 90, 92, 94, 99, 101, 127-28, 13033, 136, 139, 143-44, 146, 148, 28485, 291, 302, 315, 359, 366-67, 371, 377, 379, 383-84, 386, 388-93, 413, 458, 461, 463, 472, 477 Teichoskopie 167 Telescope see Instrument Temperament 120 Territory/Territoriality 87, 274, 395, 438, 448, 483-85, 499-500 Thaumaturgy 31, 131 Theater 44, 53, 62-63, 68-70, 75, 128, 134-37, 140, 147, 159-63, 167, 172, 223, 250, 278, 430, 494 Theater of Machines see also Theatrum machinarum 62-63, 6869, 72, 75, 153, 162-64, 168, 170 Theatricality 31, 57, 69, 71, 73-75, 125, 134, 141, 160, 500 Theatrum 72, 157, 178, 252 Theatrum Anatomicum 178 Theatrum Machinarum 152-53, 158 Theatrum Mundi see also World Theater 44, 71, 74 Theatrum Sacrum 71 Theology 26, 45, 57-58, 82, 128, 144, 195, 205, 208-10, 223, 225-28, 243, 261-63, 345, 401, 445, 450, 455, 481 Thermometer see Instrument Thought Experiment 468-69, 474 Tool 10, 16, 18, 67, 78-81, 86, 90, 112, 125, 127-28, 156, 159, 176-78, 257, 260, 263-64, 288, 297-98, 313, 348, 355, 365-67, 371, 380, 389, 395, 412, 457, 501
Index of Subjects Topography 50, 53, 56, 58, 96, 170, 183, 350 Topos/Topic 43, 58, 63, 96, 152, 172, 257, 354, 496 Trace 8, 16-17, 29, 42, 44, 48-49, 59, 81-82, 84, 110, 123, 169, 19697, 218, 278, 280, 317, 356, 369, 374, 384, 405, 431, 437, 452 Trade 65, 215, 258, 269, 274, 277, 441, 448, 467, 493 Traffic 384, 479, 493 Tragedy 70-71, 147 Travel 66, 129, 143, 214, 222-23, 242, 247, 263, 269, 273-74, 302, 315, 486 Trick 70, 110, 121, 130, 134-35, 148, 251, 288, 434, 438, 467 Trinity 48, 50, 53, 209, 239, 496 Trompe-l'œil 22-23, 141, 453 Tune 493 Universality 311, 425 Universe 73, 75, 94, 131, 134, 143, 146, 167, 208, 219, 239, 304, 372, 394 University 205, 222-28, 230, 232-33, 258, 288-89, 291-92, 354, 383, 396-97, 402, 404-09, 414-15, 488 Unrepresentability 29, 31 Uomo Universale 70, 84 Utilitarianism 79 Utopia 62, 75, 88, 290 Vacuum 14, 349, 355, 358, 467 Vanishing Point 21-22, 24-25, 27, 457, 460, 473-75 Verisimilitude 386 Vibration 258, 262, 357 Violence 72, 194, 289, 443, 450
575 Virtuality 27, 85, 130, 133, 136, 191, 260, 414, 449, 497 Virtuosity 99, 110, 125, 131, 289, 292 Visibility 6-7, 14, 16-17, 24, 26, 2829, 31-36, 40, 45, 62, 67-68, 75, 90, 95, 98, 116, 130, 132, 138, 147, 155, 158, 162, 177-78, 181-82, 20405, 209, 213, 218, 304, 319, 328, 334, 340, 348, 350, 352, 356, 36263, 367, 369, 388, 428, 460, 465, 475-76, 494 Vision see Sense Visual culture 128, 140 Visualization 4, 12, 16, 27, 40, 62, 72, 96, 98, 102, 115, 133, 180, 182-83, 193, 205, 245, 356, 459-60, 463, 474, 476 War 15-16, 88, 90, 96, 119, 143, 155, 226, 278, 297, 311, 315, 328, 397, 399, 404, 415, 442, 444-46, 493 Art of War 58 Thirty Years’ War 80, 86, 440 World War II 15, 278, 297 Weapon/Armory 88, 90, 118, 177, 498 Weight 12, 72, 95, 155, 258, 312, 317, 326-27, 441 Wisdom 168, 172, 203, 208, 413, 441, 453 Witness 25, 90, 132-33, 141, 152-53, 229, 300, 322, 402 Wonder 62-63, 65-68, 70, 129, 135, 148, 153, 168-69, 172, 239, 246, 250, 292, 339 World Theater see also Theatrum Mundi 55 World View 33 Wunderkammer 29, 31, 132, 137, 240 Zero 457, 460, 463-65, 467-77 Zoology 5, 181